research into the mechanisms of paraquat-induced multiple organ ...

research into the mechanisms of paraquat-induced multiple organ ... research into the mechanisms of paraquat-induced multiple organ ...

repositorio.aberto.up.pt
from repositorio.aberto.up.pt More from this publisher
11.12.2012 Views

� �������������������������������� ���������������������������������������� � � � � � � � � � � �� � � ������������������������������������������������������� ����������������������������������������������� ����������������������������� � ���������������������������������������� ������������

�<br />

��������������������������������<br />

����������������������������������������<br />

�<br />

�<br />

�<br />

�<br />

� �<br />

� � �<br />

� �� � �<br />

�������������������������������������������������������<br />

�����������������������������������������������<br />

�����������������������������<br />

�<br />

����������������������������������������<br />

������������


RESEARCH INTO THE MECHANISMS OF<br />

PARAQUAT- INDUCED MULTIPLE ORGAN FAILURE<br />

Development and application <strong>of</strong> antidotes and antidotal<br />

pathways for <strong>the</strong> treatment <strong>of</strong> human poisonings<br />

Ricardo Jorge Dinis Oliveira<br />

Porto, 2007<br />

FACULTAD DE FARMACIA<br />

UNIVERSIDAD DE SALAMANCA


Dissertação de candidatura ao<br />

grau de Doutor em Toxicologia apresentada<br />

à Faculdade de Farmácia da<br />

Universidade do Porto<br />

Dissertation <strong>the</strong>sis for <strong>the</strong> degree<br />

<strong>of</strong> Doctor <strong>of</strong> Philosophy in Toxicology<br />

submitted to <strong>the</strong> Faculty <strong>of</strong><br />

Pharmacy <strong>of</strong> Porto University<br />

Orientador: Pr<strong>of</strong>essor Doutor Félix Dias Carvalho (Pr<strong>of</strong>essor Associado com<br />

Agregação da Faculdade de Farmácia da Universidade do Porto;<br />

Co-orientadora: Pr<strong>of</strong>essora Doutora Maria de Lourdes Pinho de Almeida Souteiro<br />

Bastos (Pr<strong>of</strong>essora Catedrática da Faculdade de Farmácia da<br />

Universidade do Porto);<br />

Co-orientadora: Pr<strong>of</strong>essora Doutora Amparo Sánchez Navarro (Pr<strong>of</strong>essora<br />

Associada com Agregação da Faculdade de Farmácia da<br />

Universidade de Salamanca).<br />

iii


Aos meus Pais, José e Irene, e Irmã, Carla.<br />

To my Parents, José and Irene, and Sister, Carla.<br />

v


À S<strong>of</strong>ia.<br />

To S<strong>of</strong>ia.<br />

vii


viii


AUTHOR’S DECLARATION<br />

Under <strong>the</strong> terms <strong>of</strong> <strong>the</strong> Decree-Law nº 216/92, <strong>of</strong> October 13th, is hereby declared<br />

that <strong>the</strong> following original articles were prepared in <strong>the</strong> scope <strong>of</strong> this dissertation.<br />

Theoretical Background<br />

PUBLICATIONS<br />

Articles in international peer-reviewed journals<br />

I. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). Paraquat exposure as an etiological factor <strong>of</strong> Parkinson’s<br />

disease. Neurotoxicology, 27, 1110-1122.<br />

II. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2007). Paraquat poisonings: <strong>mechanisms</strong> <strong>of</strong> lung toxicity, clinical<br />

features and treatment. (Submitted for publication).<br />

Original Research<br />

I. Dinis-Oliveira, R. J., Jesus-Valle, M. J., Bastos, M. L., Carvalho, F., and Sanchez-<br />

Navarro, A. (2006). Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung. Influence <strong>of</strong> sodium<br />

depletion. Xenobiotica 36, 724-737.<br />

II. Dinis-Oliveira, R. J., Sarmento, A., Reis, P., Amaro, A., Remião, F., Bastos, M. L.,<br />

and Carvalho, F. (2006). Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival case following<br />

intake <strong>of</strong> a potential lethal dose. Pediatr Emerg Care 22, 537-540.<br />

ix


III. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). P-glycoprotein induction: an antidotal pathway for <strong>paraquat</strong><strong>induced</strong><br />

lung toxicity. Free Radic Biol Med 41, 1213-1224.<br />

IV. Dinis-Oliveira, R. J., Duarte, J. A., Remião, F., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). Single high dose dexamethasone treatment decreases <strong>the</strong><br />

pathological effects and increases <strong>the</strong> survival rat <strong>of</strong> <strong>paraquat</strong>-<strong>into</strong>xicated rats.<br />

Toxicology 227, 73-85.<br />

V. Dinis-Oliveira, R. J., Sousa, C., Remião, F., Duarte, J. A., Sanchez-Navarro, A.,<br />

Bastos, M. L., and Carvalho, F. (2007). Full survival <strong>of</strong> <strong>paraquat</strong>-exposed rats after<br />

treatment with sodium salicylate. Free Radic Biol Med 42, 1017-1028.<br />

VI. Dinis-Oliveira, R. J., Sousa, C., Remião, F., Duarte, J. A., Sanchez-Navarro, A.,<br />

Bastos, M. L., and Carvalho, F. (2007). Effects <strong>of</strong> sodium salicylate in <strong>the</strong> <strong>paraquat</strong><strong>induced</strong><br />

apoptotic events in rat lungs. Free Radic Biol Med 43, 48-61.<br />

Original Research<br />

x<br />

Abstracts in international peer-reviewed journals<br />

I. Dinis-Oliveira, R. J., Jesus-Valle, M. J., Bastos, M. L., Carvalho, F., and Sanchez-<br />

Navarro, A. (2005). Influence <strong>of</strong> sodium depletion on <strong>the</strong> kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong><br />

isolated rat lung. Toxicol Lett 158 (Suppl. 1): S213.<br />

II. Dinis-Oliveira, R. J., Sarmento, A., Reis, P., Amaro, A., Remião, F., Bastos, M. L.,<br />

and Carvalho, F. (2005). Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival case following<br />

intake <strong>of</strong> a potential lethal dose. Toxicol Lett 15 (Suppl. 1): S241.<br />

III. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). A new and vital antidotal pathway for <strong>paraquat</strong> poisonings<br />

more than 60 years later: Induction <strong>of</strong> lung P-glycoprotein. Toxicol Lett 164 (Suppl. 1):<br />

S75.


IV. Dinis-Oliveira, R. J., Duarte, J. A., Remião, F., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). Dexamethasone treatment decreases <strong>the</strong> pathological effects<br />

and increases <strong>the</strong> survival rate <strong>of</strong> <strong>paraquat</strong>-<strong>into</strong>xicated rats. Toxicol Lett 164 (Suppl. 1):<br />

S237-S238.<br />

V. Dinis-Oliveira, R. J., Remião, F., Sanchez-Navarro, A., Duarte, J. A., Bastos, M. L.,<br />

and Carvalho, F. (2007). Recent developments in <strong>the</strong> <strong>the</strong>rapy <strong>of</strong> <strong>paraquat</strong> poisoning.<br />

Clin Toxicol (Phila) 45:376 (invited keynote presentation).<br />

Original Research<br />

Patents<br />

I. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. The use <strong>of</strong> <strong>the</strong> process <strong>of</strong> induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> Pglycoprotein<br />

(P-gp) in <strong>the</strong> treatment <strong>of</strong> xenobiotic-<strong>induced</strong> <strong>into</strong>xications in mammals.<br />

Portuguese Patent Nº 103420/06 (2006) and International Patent Nº<br />

PCT/IB2007/050144 (2007).<br />

II. Dinis-Oliveira, R. J., Remião, F., Duarte, J. A., Sanchez-Navarro, A., Bastos, M. L.,<br />

and Carvalho, F. (2006). The use <strong>of</strong> salicylate as an antidote <strong>of</strong> <strong>paraquat</strong> <strong>into</strong>xications in<br />

humans. Portuguese Patent Nº 103480/06 (2006) and International Patent<br />

Nº PCT/IB2007/051799 (2007).<br />

Under <strong>the</strong> terms <strong>of</strong> <strong>the</strong> referred Decree-Law, <strong>the</strong> author declares that he afforded<br />

a major contribution to <strong>the</strong> conceptual design and technical execution <strong>of</strong> <strong>the</strong> work,<br />

interpretation <strong>of</strong> <strong>the</strong> results and manuscript preparation <strong>of</strong> <strong>the</strong> published articles<br />

included in this dissertation.<br />

xi


xii<br />

Programa Operacional da Ciência e Inovação 2010<br />

MINISTÉRIO DA CIÊNCIA, INOVAÇÃO E ENSINO SUPERIOR<br />

The candidate performed <strong>the</strong> experimental work with a doctoral fellowship<br />

(SFRH/BD/13707/2003) supported by <strong>the</strong> “Fundação para a Ciência e a Tecnologia”,<br />

which also participate with grants to attend in international meetings and for <strong>the</strong><br />

graphical execution <strong>of</strong> this <strong>the</strong>sis. The Faculty <strong>of</strong> Pharmacy <strong>of</strong> <strong>the</strong> University <strong>of</strong> Porto<br />

(Portugal) and <strong>of</strong> Salamanca (Spain) provided <strong>the</strong> facilities and logistical supports.


ACKNOWLEDGMENTS<br />

There are lots <strong>of</strong> people I would like to thank for a huge variety <strong>of</strong> reasons. Despite<br />

this dissertation represents an individual work, it involved more than three years <strong>of</strong><br />

cooperation between many people. It is a pleasure to express my truthful gratitude to all<br />

who made this <strong>the</strong>sis possible with words <strong>of</strong> encouragement. This is perhaps <strong>the</strong> easiest<br />

and hardest chapter that I have to write. It will be simple to name all <strong>the</strong> people that<br />

helped to get this done, but it will be tough to thank <strong>the</strong>m enough. I will none<strong>the</strong>less<br />

try…<br />

Firstly and at HEAD, I would like to thank my Supervisor, Pr<strong>of</strong>. Félix Dias<br />

Carvalho for his guidance and support during <strong>the</strong> course <strong>of</strong> this work. I could not have<br />

imagined having a better advisor and mentor for my PhD, and without his commonsense,<br />

knowledge, and perceptiveness I would never have finished. It is difficult to<br />

overstate my gratitude. With his enthusiasm, his inspiration, and his great efforts to<br />

explain things clearly and simply, he helped me to make <strong>the</strong> investigation fun for me.<br />

Throughout my <strong>the</strong>sis period, he provided encouragement, sound advice, good teaching,<br />

good company and friendship, and lots <strong>of</strong> good ideas. I would have been lost without<br />

him. It has been a privilege to be able to learn from his example as a <strong>research</strong>er and<br />

mentor. I have been extremely lucky to have a supervisor who cared so much about my<br />

work, and who responded to my questions and queries so promptly. Definitely, you are<br />

an example to follow. His careful support was highly constructive for my growth and<br />

makes him a true friend.<br />

I would like to extend my gratitude to Pr<strong>of</strong>. Maria de Lourdes Bastos, my cosupervisor,<br />

for her help and for much type <strong>of</strong> facilities she provided me during all <strong>the</strong>se<br />

approximately three years. Her dynamic and prompt way <strong>of</strong> being and her work<br />

capacity are remarkable characteristics to pursue. With all respect, her “youth” <strong>of</strong> spirit<br />

is an example <strong>of</strong> life.<br />

I am very grateful to Pr<strong>of</strong>essor Amparo Sánchez Navarro, my co-supervisor, for<br />

giving me <strong>the</strong> opportunity to work at Department <strong>of</strong> Pharmacy and Pharmaceutical<br />

xiii


Technology, Faculty <strong>of</strong> Pharmacy <strong>of</strong> University <strong>of</strong> Salamanca, for her welcoming<br />

during my stays in Salamanca, valuable guidance, numerous critical appraisals and<br />

scientific discussions, not always easy because <strong>of</strong> <strong>the</strong> distance. The answers to my email<br />

questions and <strong>the</strong> comments on my manuscripts were always sent back sooner than<br />

I expected.<br />

I'm grateful to my “supervisor” Pr<strong>of</strong>essor José Alberto Duarte from CIAFEL <strong>of</strong><br />

<strong>the</strong> Faculty <strong>of</strong> Sport <strong>of</strong> <strong>the</strong> University <strong>of</strong> Porto. Our discussions throughout <strong>the</strong> nights<br />

would certainly be important to my constant academic growth. His arduous, rigorous,<br />

meticulous work and valuable help in many steps was <strong>of</strong> incontestable value. I am<br />

deeply indebted to Pr<strong>of</strong>essor José Alberto Duarte for his involvement, unlimited<br />

contribution to this study, for <strong>the</strong> many insightful conversations during <strong>the</strong> development<br />

<strong>of</strong> new ideas, and for helpful comments on <strong>the</strong> text. I want to express that he was an<br />

outstanding advisor and an excellent pr<strong>of</strong>essor. I could always count with him, and I<br />

will always be grateful for his expertise and support. When I met him for <strong>the</strong> first time I<br />

was far from realise that it would certainly be very difficult to conclude this <strong>the</strong>sis<br />

without his friendship. Thanks for all Pr<strong>of</strong>essor.<br />

I’m grateful to Pr<strong>of</strong>. Fernando Remião for his assistance in helping to supervise<br />

me, providing resources and subjects, <strong>of</strong>fering direction and penetrating criticism, for<br />

many insightful conversations during <strong>the</strong> development <strong>of</strong> new ideas in this <strong>the</strong>sis, for<br />

<strong>the</strong> fun environment he provided, and for making <strong>the</strong> COHiTEC project possible. I’m<br />

also thankful to have benefited from his previous works on <strong>paraquat</strong> quantification and<br />

for all <strong>the</strong> remarkable worries he demonstrated in supplying good laboratory conditions.<br />

Pr<strong>of</strong>essor, I am grateful in every possible way and hope to keep up our collaboration in<br />

<strong>the</strong> future with our entrepreneur life style.<br />

I acknowledge “Fundação para a Ciência e a Tecnologia” for my doctoral<br />

fellowship (SFRH / BD / 13707 / 2003) and for <strong>the</strong> financial support <strong>of</strong> this dissertation.<br />

Thanks for <strong>the</strong> contribution.<br />

I acknowledge “REQUIMTE”, associated laboratory, for <strong>the</strong> financial support <strong>of</strong><br />

<strong>the</strong> laboratory work.<br />

xiv


To Engineer Elisa, <strong>the</strong> Mum <strong>of</strong> <strong>the</strong> department, besides <strong>the</strong> important life lessons<br />

she gave me, I’m thankful for her friendship and companionship forever.<br />

To all my toxicology department friends, I would like to thank <strong>the</strong>ir laboratorial<br />

and fieldwork assistance as well as <strong>the</strong>ir companionship and <strong>the</strong> permanent support and<br />

motivation. In particular, my appreciation to Carla, Cecília, Estela, Helena Carmo,<br />

Helena Pontes, João, Renata, Teresa e Vera, for <strong>the</strong>ir friendship and fun environment in<br />

which I learnt, grew and for <strong>the</strong> helpful support along <strong>the</strong>se years. A special encourage<br />

word to Carla for her irreplaceable and exceptional support during all laboratorial<br />

assays. She really has an appreciable work capacity in <strong>the</strong> laboratory. To Görkem<br />

Mergen for all <strong>the</strong> entertaining moments I send you a big hug.<br />

To Mrs. Julia, Graziela and Conceição for <strong>the</strong>ir prompt support whenever<br />

necessary in solving technical problems, for <strong>the</strong> careful washing and handling with <strong>the</strong><br />

laboratory material and for <strong>the</strong>ir efficient resolution <strong>of</strong> logistic problems. We all miss<br />

Mrs. Graziela very much in <strong>the</strong> Faculty. Wherever you are now, we hope you are in<br />

peace... Thanks for all <strong>the</strong> help.<br />

To Mrs. Casemira for her prompt, gentleness and sympathy in provide missing<br />

laboratory material to my work persecution.<br />

I would like to thank to my high school chemistry teacher (Pr<strong>of</strong>essor Augusta<br />

Silveira) for <strong>the</strong> kind assistance, advices and help in conferring some justice to <strong>the</strong><br />

marks and consequently to make everything possible. A special thank for that Pr<strong>of</strong>essor.<br />

A special thank to David Costa for all <strong>the</strong> help and knowledge he shared without<br />

asking anything to exchange. You are a good guy.<br />

I’m gratified to Pr<strong>of</strong>essor Eduarda Fernandes for her collaboration in <strong>the</strong><br />

persecution <strong>of</strong> <strong>the</strong> work, mainly helping me leading with <strong>the</strong> Nuclear magnetic<br />

resonance (NMR) spectrums.<br />

I’m grateful to Pr<strong>of</strong>essor Franklim Marques for had provided <strong>the</strong> facilities in <strong>the</strong><br />

beginning <strong>of</strong> <strong>the</strong> <strong>research</strong>.<br />

xv


A special word <strong>of</strong> acknowledge to Pr<strong>of</strong>essor Natércia, to Pr<strong>of</strong>essor Georgina and<br />

to Margarida for <strong>the</strong>ir collaboration leading with <strong>the</strong> TUNEL experiments.<br />

I’m thankful to Ana Margarida for her amiable support regarding to animals’ care<br />

and treating protocols.<br />

To Rita Ferreira from CIAFEL who participated and well contributed in several<br />

biochemical quantifications.<br />

To Mrs. Celeste Resende for her technical assistance regarding <strong>the</strong> samples<br />

processing for Light and Electron Microscopy. Naturally, I apologise for <strong>the</strong> time I still<br />

you.<br />

Doctor Cândida, <strong>the</strong> librarian, I am grateful to her assistance in providing <strong>the</strong><br />

bibliography and for that she deserve special mention.<br />

To all my friends from Department <strong>of</strong> Pharmacy and Pharmaceutical Technology<br />

<strong>of</strong> <strong>the</strong> Faculty <strong>of</strong> Pharmacy <strong>of</strong> <strong>the</strong> University <strong>of</strong> Salamanca for his friendship in all<br />

moments we shared toge<strong>the</strong>r. A special thank for <strong>the</strong> beginning <strong>of</strong> my scientific<br />

edification.<br />

I thankfully acknowledge <strong>the</strong> promptness help and friendly contribution <strong>of</strong> my<br />

friends Ricardo Silvestre, Joana Maciel and Joana Tavares from <strong>the</strong> Department <strong>of</strong><br />

Biochemistry.<br />

To S<strong>of</strong>ia, my girl-friend, for her sincerely friendship, constant care in good and<br />

bad moments, emotional support, camaraderie, entertainment, and caring she provided. I<br />

would like to wish you <strong>the</strong> same you desire for me. She was my own "soul out <strong>of</strong> my<br />

soul," who kept my spirits up when <strong>the</strong> muses failed me.<br />

To my lovely Mum, Maria Irene, and Dad, José Carneiro, and to my sister, Carla<br />

Isabel, who experienced all <strong>of</strong> <strong>the</strong> ups and downs <strong>of</strong> my <strong>research</strong> and for all <strong>the</strong><br />

patience, to <strong>the</strong>m I dedicate this <strong>the</strong>sis. Despite far away most <strong>of</strong> <strong>the</strong> times I know that<br />

we are always close. You are unique and I’m very proud for being your sun and bro<strong>the</strong>r.<br />

xvi


To my godfa<strong>the</strong>r, José P<strong>into</strong>, and to Paula and Tó for all <strong>the</strong> advices in anxiety and<br />

apprehension moments, my special acknowledge.<br />

To finalize I want to give a word for <strong>the</strong> loving memory <strong>of</strong> my godmo<strong>the</strong>r and<br />

grandparents, Joaquim Dinis and Severino da Costa. I miss you so much…<br />

Is that everyone?<br />

Ricardo Dinis, Porto, 2007<br />

xvii


xviii


ABSTRACT<br />

Paraquat (PQ) is a popular herbicide and it is probably one <strong>of</strong> <strong>the</strong> most studied<br />

pesticides. This interest is related not only to its relevance as a human poison, but also<br />

to its specific toxicological <strong>mechanisms</strong>, namely in <strong>the</strong> area <strong>of</strong> oxidative stress, making<br />

it an excellent <strong>research</strong> tool. PQ has a proven safety record when used properly for its<br />

intended purpose. Its safety in use can, at least in part, be explained by its lack <strong>of</strong><br />

absorption ei<strong>the</strong>r by inhalation or through <strong>the</strong> intact skin. This is because <strong>the</strong> spray<br />

droplets generated by agricultural equipment are too large in diameter (>5 μm) to be<br />

inhaled and because <strong>the</strong> skin provides an effective, impermeable barrier to <strong>the</strong><br />

absorption <strong>of</strong> PQ. However, over <strong>the</strong> past 44 years, <strong>the</strong>re have been numerous fatalities<br />

following accidental or deliberate ingestion <strong>of</strong> this weed-killer. Poisonings have<br />

received considerable attention in <strong>the</strong> medical, scientific, and popular press. The<br />

<strong>into</strong>xication cases have <strong>of</strong>ten been dramatic because <strong>of</strong> <strong>the</strong> protracted and inexorable<br />

course <strong>of</strong> <strong>the</strong> illness and <strong>the</strong> absence <strong>of</strong> any effective antidote. Fatality rates have been<br />

over 50%. In addition, very little pre-clinical <strong>research</strong> has been successful in applying<br />

antidotes that shown to work in animal studies <strong>into</strong> clinical practice. PQ accumulates<br />

mainly in <strong>the</strong> lung (pulmonary concentrations can be six to ten times higher than those<br />

in <strong>the</strong> plasma), where it is retained even when blood levels start to decrease. The<br />

pulmonary effects can be readily explained by <strong>the</strong> participation <strong>of</strong> <strong>the</strong> polyamine<br />

transport system abundantly expressed in <strong>the</strong> membrane <strong>of</strong> alveolar cells type I, II and<br />

Clara cells. Fur<strong>the</strong>r downstream at <strong>the</strong> toxicodynamic level, <strong>the</strong> molecular mechanism<br />

<strong>of</strong> PQ toxicity is based on redox cycling and intracellular oxidative stress generation.<br />

During <strong>the</strong> last years, our <strong>research</strong> group has been a reference in <strong>the</strong> field <strong>of</strong> PQ<br />

toxicity to hospitals in <strong>the</strong> centre and north <strong>of</strong> Portugal. According, this dissertation was<br />

primarily aimed to describe a successful clinical case, regarding <strong>the</strong> <strong>into</strong>xication <strong>of</strong> a 15year-old<br />

girl by a presumed lethal dose <strong>of</strong> PQ. Besides <strong>the</strong> measures for decreasing PQ<br />

absorption and increasing its elimination, o<strong>the</strong>r protective procedures were applied<br />

aiming to reduce <strong>the</strong> production <strong>of</strong> reactive oxygen species (ROS), scavenge and repair<br />

ROS-<strong>induced</strong> lesions, and to reduce inflammation. The status-<strong>of</strong>-<strong>the</strong>-art concerning <strong>the</strong><br />

biochemical and toxicological aspects <strong>of</strong> PQ poisoning and <strong>the</strong> pharmacological basis <strong>of</strong><br />

<strong>the</strong> respective treatment protocol was presented. It was conducted an intensive and<br />

xix


aggressive treatment, based on <strong>the</strong> high urinary and plasmatic PQ concentrations, and<br />

accordingly to <strong>the</strong> positive outcome, <strong>the</strong> <strong>the</strong>rapeutic protocol followed could be a<br />

promising treatment <strong>of</strong> PQ human <strong>into</strong>xications. Besides, were also good prognostic<br />

factors, young age, lesser degrees <strong>of</strong> leukocytosis and acidosis, and <strong>the</strong> absence <strong>of</strong> renal,<br />

hepatic, and pancreatic failures on admission after acute PQ poisoning.<br />

xx<br />

In this <strong>the</strong>sis, <strong>the</strong> usefulness <strong>of</strong> <strong>the</strong> isolated rat lung was firstly explored. Such<br />

model was applied to characterize <strong>the</strong> toxicokinetic behaviour <strong>of</strong> PQ in this tissue after<br />

bolus injection under standard experimental conditions as well as to evaluate <strong>the</strong><br />

influence <strong>of</strong> iso-osmotic replacement <strong>of</strong> sodium by lithium in <strong>the</strong> perfusion medium.<br />

The obtained results showed that <strong>the</strong> isolated rat lung model is a useful technique for<br />

PQ toxicokinetic studies. It was also observed that sodium-depletion in <strong>the</strong> perfusion<br />

medium leads to a decreased uptake <strong>of</strong> PQ in <strong>the</strong> isolated rat lung although it seems that<br />

this condition does not contribute to improve <strong>the</strong> elimination <strong>of</strong> PQ once <strong>the</strong> herbicide<br />

reaches <strong>the</strong> extravascular structures <strong>of</strong> <strong>the</strong> tissue. In spite <strong>of</strong>, techniques <strong>of</strong> tissue<br />

isolation and perfusion <strong>of</strong>fer an excellent alternative to characterize <strong>the</strong> kinetic pr<strong>of</strong>ile<br />

for a tissue in a single animal and avoids <strong>the</strong> inter-individual variability in each single<br />

curve, leading as well to a corresponding reduction in curve replicates and hence a<br />

substantial reduction in <strong>the</strong> number <strong>of</strong> animals used (5-8 versus 50-80/tissue), soon we<br />

notice that PQ toxicity is a myriad <strong>of</strong> factors, toge<strong>the</strong>r contributing to a death outcome<br />

and it would be better studied by in vivo approaches.<br />

According to this desideratum, secondly it is described a procedure, through <strong>the</strong><br />

induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-glycoprotein by <strong>the</strong> administration <strong>of</strong> a single high<br />

dose <strong>of</strong> dexamethasone (DEX) to Wistar rats, that leads to a remarkable decrease <strong>of</strong> PQ<br />

accumulation in <strong>the</strong> lung, toge<strong>the</strong>r with an increase <strong>of</strong> its faecal excretion and a<br />

subsequent decrease <strong>of</strong> several biochemical and histopathological biomarkers <strong>of</strong><br />

toxicity. The obtained results shown that DEX also ameliorated <strong>the</strong> biochemical and<br />

histological liver alterations <strong>induced</strong> by PQ in Wistar rats. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong>se<br />

improvements were not observed in kidney and spleen <strong>of</strong> DEX treated rats. The sum <strong>of</strong><br />

<strong>the</strong>se effects was clearly positive, since it was observed an increased survival rate,<br />

which indicates that high dosage DEX treatment constitutes an important and valuable<br />

<strong>the</strong>rapeutic tool to be used against PQ-<strong>induced</strong> toxicity.<br />

Finally <strong>the</strong> role <strong>of</strong> <strong>the</strong> apoptosis, oxidative stress, platelet aggregation, nuclear<br />

factor (NF)-κB activation and fibrosis in PQ-<strong>induced</strong> lung toxicity, as well as <strong>the</strong><br />

remarkable healing effects obtained by <strong>the</strong> administration <strong>of</strong> sodium salicylate (NaSAL,


200 mg/Kg i.p.), were assessed. The obtained results exceeded our best expectations<br />

since not only <strong>the</strong> toxicity was reverted but, most significantly, full survival <strong>of</strong> <strong>the</strong> PQ-<br />

<strong>into</strong>xicated rats treated with NaSAL was observed. It may be postulated that NaSAL is<br />

<strong>the</strong> first real PQ antidote described with such degree <strong>of</strong> success.<br />

Of note, <strong>the</strong> administrations <strong>of</strong> DEX and NaSAL were given two hours after<br />

<strong>into</strong>xication <strong>of</strong> rats with PQ, a lag time that confers realism to be applied in humans,<br />

since this chronological time corresponds to longer biological time for humans and<br />

<strong>the</strong>refore this represent <strong>the</strong> probable time that passes between <strong>the</strong> herbicide ingestion<br />

and <strong>the</strong> begin <strong>of</strong> <strong>the</strong> medical cares.<br />

In conclusion, <strong>the</strong> results <strong>of</strong> this dissertation suggest that high doses <strong>of</strong> DEX<br />

and/or NaSAL are <strong>the</strong>rapeutic approaches with potential to be applied in humans,<br />

though, fur<strong>the</strong>r pre-clinical studies are needed particularly those aimed to explain in<br />

more detail <strong>the</strong> mode <strong>of</strong> action <strong>of</strong> <strong>the</strong>se interesting drugs in <strong>the</strong> protection against PQ<strong>induced</strong><br />

lung damage.<br />

xxi


xxii


RESUMO<br />

O <strong>paraquat</strong>o (PQ) é provavelmente um dos pesticidas mais estudados. Este<br />

interesse não está apenas relacionado com a sua relevância como um xenobiótico<br />

envolvido em <strong>into</strong>xicações humanas, mas também devido aos seus mecanismos<br />

toxicológicos, nomeadamente na área do stress oxidativo onde representa uma<br />

ferramenta de extrema importância. O PQ tem uma comprovada segurança quando<br />

usado devidamente, podendo esta ser em parte explicada pela ausência de absorção por<br />

via inalatória e através da pele integra. Isto porque, o tamanho das gotículas produzidas<br />

pelos equipamentos agrícolas, apresentam um diâmetro demasiadamente grande (>5<br />

μm) para serem inaladas e também porque a pele constitui uma barreira impermeável à<br />

absorção do PQ. No entanto, nos últimos 44 anos, o PQ tem sido a causa de diversas<br />

mortes sobretudo por ingestão, acidental ou voluntária, mas também por exposição<br />

dérmica. As <strong>into</strong>xicações rapidamente receberam considerável atenção da comunidade<br />

médica, científica e dos meios de comunicação social. Apesar das razões para o uso<br />

deste herbicida como agente de suicídio sejam difíceis de determinar, o principal<br />

responsável parece ser estes últimos que informam e documentam os suicídios<br />

resultantes da toxicidade aguda do PQ.<br />

Os casos de <strong>into</strong>xicação têm sido dramáticos, muito porque o curso da doença é<br />

rápido e pela total ausência de um antídoto ou tratamento eficaz, dependendo a<br />

sobrevivência dos <strong>into</strong>xicados da quantidade ingerida e do tempo que decorre até o<br />

início das intervenções médicas para eliminar o PQ que ainda não foi absorvido e/ou<br />

captado pelas células. As taxas de letalidade cifram-se em valores acima dos 50%. Além<br />

do mais, muito pouco do resultante da investigação pré-clínica tem sido aplicado com<br />

sucesso na apática clínica. O PQ acumula-se maioritariamente no pulmão, onde as<br />

concentrações podem atingir seis ou mesmo dez vezes as concentrações plasmáticas,<br />

ficando aí mesmo quando os níveis plasmáticos começam a diminuir. Os efeitos<br />

pulmonares podem ser facilmente explicados pela participação do transportador das<br />

poliaminas endógenas abundantemente expressado na membrana os pneumócitos Tipo<br />

I, II, e nas células Clara. Ao nível toxicodinâmico, o mecanismo de toxicidade do PQ é<br />

baseado no ciclo redox e na geração de espécies reactivas do oxigénio (ROS).<br />

xxiii


Durante os últimos anos, o nosso grupo de investigação tem sido uma referência<br />

para hospitais no centro e norte de Portugal no que se refere à toxicidade do PQ. Em<br />

conformidade, o primeiro objectivo desta dissertação foi descrever um caso clínico de<br />

sucesso, de uma jovem de 15 anos de idade que voluntariamente ingeriu uma dose letal<br />

de PQ. Para além de medidas destinadas a diminuir a absorção do PQ ou aumentar a sua<br />

eliminação, outras medidas protectoras foram também seguidas de modo a diminuir a<br />

produção das ROS, captar as ROS, reparação das lesões produzidas pelas ROS e<br />

diminuir a inflamação. O estado da arte referente aos aspectos bioquímicos e<br />

toxicológicos do PQ e a base farmacológica do protocolo terapêutico seguido foi<br />

apresentado. Conduziu-se um tratamento intensivo e agressivo baseado nos altos níveis<br />

urinários e plasmáticos e tendo em conta o resultado positivo, foi possível concluir que<br />

o protocolo seguido poderá ser prometedor no tratamento das <strong>into</strong>xicações humanas<br />

pelo PQ. Foram também factores de prognósticos favoráveis, a juventude, menores<br />

graus de leucocitose, ausência de falha renal, hepática e pancreática à admissão após<br />

<strong>into</strong>xicação aguda.<br />

De forma a reduzir a morbilidade e mortalidade associada às <strong>into</strong>xicações pelo<br />

PQ, primeiramente nesta dissertação, explorou-se a utilidade do modelo de pulmão<br />

isolado de rato com o objectivo de caracterizar o comportamento toxicocinético do PQ<br />

neste tecido, após injecção por bólus sob condições padrão assim como para avaliar a<br />

influência da substituição iso-osmótica do sódio pelo lítio do meio de perfusão. Os<br />

resultados obtidos comprovaram a aplicabilidade do modelo de pulmão isolado no<br />

estudo da toxicocinética do PQ. Observou-se também que a depleção de sódio do meio<br />

de perfusão conduziu a diminuição da captação pulmonar de PQ neste modelo de<br />

estudo, apesar desta condição não levar a um aumento da eliminação do PQ deste<br />

tecido, uma vez alcançado as estruturas extracelulares. Apesar das técnicas de<br />

isolamento e perfusão de órgãos <strong>of</strong>erecerem uma excelente alternativa para a<br />

caracterização do perfil toxicocinético em um tecido de um animal, evitando as<br />

diferenças de variabilidade interindividuais em cada curva, conduzindo também a uma<br />

redução dos replicados da curva e como tal uma redução do numero de animais usados<br />

(5-8 versus 50-80/tecido), rapidamente se constatou que a toxicidade do PQ resulta de<br />

uma miríade de factores, juntos contribuindo para morte, os quais seriam melhor<br />

estudados usando uma abordagem in vivo.<br />

De acordo com este desiderato, secundariamente nesta dissertação é descrito um<br />

procedimento, através da indução da síntese de novo da glicoproteína-P (P-gp) por<br />

xxiv


administração de uma dose única mas elevada de dexametasona (DEX) a ratos Wistar, o<br />

qual conduziu a uma marcada diminuição da acumulação pulmonar do PQ para menos<br />

de 40% em apenas 24 horas, associado a um aumento da sua excreção fecal e<br />

consequente melhoramento de vários parâmetros bioquímicos e histopatológicos de<br />

toxicidade. Os resultados obtidos demonstraram que o tratamento com DEX também<br />

produziu uma melhoria nas alterações dos mesmos parâmetros no fígado de animais<br />

<strong>into</strong>xicados pelo PQ. Apesar de tais resultados positivos não terem sido registados no<br />

baço e no rim após tratamento com DEX, o somatório destes efeitos foi claramente<br />

benéfico, uma vez que foi observado um aumento da percentagem de sobrevivência<br />

para 50% ao final de 10 dias em oposição aos 10% reportados em estudos anteriores<br />

realizados por outros grupos de investigação e utilizando múltiplas abordagens<br />

terapêuticas. Pode-se desta forma concluir que altas doses de DEX constituem uma<br />

importante e valiosa ferramenta terapêutica a ser usada nas <strong>into</strong>xicações pelo PQ.<br />

Apesar da extrema importância deste estudo, a persistente lacuna relacionada com<br />

a inexistência de um antídoto que garanta 100% de sobrevivência das <strong>into</strong>xicações pelo<br />

PQ motivou finalmente o último estudo documentado nesta dissertação, no qual se<br />

demonstrou que o salicilato de sódio (NaSAL, 200 mg/Kg i.p.) tem um elevado<br />

potencial para constituir um verdadeiro antídoto das <strong>into</strong>xicações pelo PQ. Neste estudo<br />

foi avaliado o papel da apoptose, do stress oxidativo, da activação plaquetária, do factor<br />

de transcrição pro-inflamatório, Nuclear Factor (NF)-κB e da fibrose, assim como os<br />

remarcáveis efeitos protectores do NaSAL na toxicidade pulmonar induzida pelo PQ.<br />

Verificou-se que a alteração destes factores como consequência da exposição ao PQ,<br />

pode ser significativamente inibida pelo NaSAL com a consequente recuperação dos<br />

animais <strong>into</strong>xicados. Na verdade, este estudo revelou-se muito mais importante do que à<br />

partida era esperado, pois pela primeira vez, foi demonstrado que com este fármaco e<br />

com apenas uma dose foi possível reverter toda a toxicidade dos animais <strong>into</strong>xicados<br />

com o PQ e garantir uma percentagem de sobrevivência de 100%.<br />

De salientar que o efeito antidotal da DEX e do NaSAL resultou da administração<br />

destes fármacos 2 horas após a <strong>into</strong>xicação dos animais com o PQ. Este intervalo<br />

confere um maior realismo na aplicação às <strong>into</strong>xicações humanas, pois reflecte, em<br />

grande parte dos casos, o tempo que medeia entre a ingestão deste herbicida e o início<br />

do possível tratamento hospitalar do paciente.<br />

Em conclusão, os resultados desta dissertação sugerem que altas doses de DEX<br />

e/ou de NaSAL constituem importantes abordagens terapêuticas com potencial<br />

xxv


aplicativo em humanos, apesar de mais estudos pré-clínicos serem necessários,<br />

nomeadamente aqueles destinados a clarificar modo de acção do NaSAL na prevenção<br />

da lesão pulmonar originada pelo PQ e também aqueles com o objectivo de encontrar<br />

novos, específicos e mais potentes indutores da síntese de novo da P-gp, isentos dos<br />

efeitos adversos que são bem conhecidos para os glucocorticóides.<br />

xxvi


Ab, antibody;<br />

ACE, angiotensin-converting enzyme;<br />

AP-1, activator protein-1;<br />

ARDS, acute respiratory distress syndrome;<br />

ATP, adenosine triphosphate;<br />

BALF, bronchoalveolar lavage fluid;<br />

BHs, bipyridylium herbicides;<br />

BW, body weight;<br />

CHP, charcoal hemoperfusion;<br />

CLCr, creatinine clearance;<br />

CLPQ, <strong>paraquat</strong> clearance;<br />

Cmax, maximum plasma concentration;<br />

CNS, central nervous system;<br />

CP, cyclophosphamide;<br />

CP51, hydroxypyridin-4-one;<br />

Cyt c, Cytochrome c;<br />

DEX, dexamethasone;<br />

DFO, desferoxamine;<br />

ABBREVIATIONS LIST<br />

DLCO, lung carbon monoxide diffusing capacity;<br />

DNA, deoxyribonucleic acid;<br />

ERG, electroretinogram;<br />

Fe 2+ , ferrous ion;<br />

Fe 3+ , ferric ion;<br />

FiO2, fraction <strong>of</strong> inspired oxygen;<br />

FR, Fenton reaction;<br />

FRD, ferrodoxin;<br />

G6PD, glucose-6-phosphate dehydrogenase;<br />

GFR, glomerular filtration rate;<br />

GGO, ground-glass opacification;<br />

GIT, gastrointestinal tract;<br />

xxvii


GPx, glutathione peroxidase;<br />

Gred, glutathione reductase;<br />

GSH, reduced glutathione;<br />

GSSG, oxidized glutathione;<br />

H2O2, hydrogen peroxide;<br />

HMP, hexose monophosphate pathway;<br />

HO . , hydroxyl radical;<br />

HRCT, high-resolution computed tomography;<br />

HWR, Haber-Weiss reaction;<br />

i.p., intraperitoneal;<br />

i.v., intravenous;<br />

ICI, Imperial Chemical Industries (now Syngenta);<br />

Km, Michaelis-Menten constant;<br />

LPO, lipid peroxidation;<br />

MDA, malondialdehyde;<br />

MINA, 4-methylisonicotinic acid;<br />

MGBG, methylglyoxal bis-(guanylhydrazone);<br />

MP, methylprednisolone;<br />

Na + , sodium;<br />

NAC, N-acetylcysteine;<br />

NADP + , oxidized nicotinamide adenine dinucleotide phosphate;<br />

NADPH, reduced nicotinamide adenine dinucleotide phosphate;<br />

NaSAL, sodium salicylate;<br />

NF-κB, nuclear factor kappa-B;<br />

NMN, N-methylnicotinamide;<br />

NO, nitric oxide;<br />

NOS, nitric oxide synthase;<br />

O2, oxygen;<br />

O2 .- , superoxide radical;<br />

PaCO2, partial pressure <strong>of</strong> carbon dioxide in arterial blood;<br />

PAH, p-aminohippurate;<br />

PaO2, partial pressure <strong>of</strong> oxygen in arterial blood;<br />

PAO2, partial pressure <strong>of</strong> oxygen in <strong>the</strong> alveolus;<br />

PEEP, positive end-expiratory pressure;<br />

xxviii


PFTs, pulmonary function tests;<br />

P-gp, P-glycoprotein;<br />

PQ or PQ 2+ , <strong>paraquat</strong>;<br />

PQ •+ , <strong>paraquat</strong> monocation free radical;<br />

PUS, polyamine uptake system;<br />

RNA, ribonucleic acid;<br />

ROS, reactive oxygen species;<br />

s.c., subcutaneous;<br />

SH, thiol;<br />

SOD, superoxide dismutase;<br />

t1/2, half-life;<br />

Tmax, time to maximum plasma concentration;<br />

TPC, TUNEL-positive cells;<br />

TUNEL, terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate<br />

nick end-labeling;<br />

VER, verapamil;<br />

Vmax, maximal rate;<br />

WBC, white blood cell;<br />

XD, xanthine dehydrogenase;<br />

XO, xanthine oxidase.<br />

xxix


xxx


INDEX OF FIGURES<br />

Fig. 1 – Chemical structures <strong>of</strong> <strong>paraquat</strong> (PQ) and diquat (DQ)...................................... 3<br />

Fig. 2 – Equilibrium dynamics for <strong>paraquat</strong> between <strong>the</strong> soil and soil solution. Adapted<br />

from Roberts et al. (Roberts et al., 2002)......................................................................... 6<br />

Fig. 3 – Herbicidal mechanism <strong>of</strong> <strong>paraquat</strong>. In photosystem I (PS I), plastocyanin (PC)<br />

transfers its electron (e - ) through a series <strong>of</strong> steps (P700, A0, A1, FX, FA) to ferrodoxin<br />

(FRD) and finally to NADP + . PQ ion (PQ 2+ ) binds near <strong>the</strong> FRD binding site in PS I and<br />

accepts an e - , becoming <strong>paraquat</strong> monocation free radical (PQ •+ ), which initiates a series<br />

<strong>of</strong> reactions leading to cell membrane disruption and plant death. The formation <strong>of</strong> such<br />

free radicals stops electron transport to NADP + and effectively inhibits normal<br />

functioning <strong>of</strong> PS I. A, Ferrodoxin-NADP + reductase. .................................................... 8<br />

Fig. 4 – Photochemical and microbial degradation <strong>of</strong> PQ. Adapted from Slade (Slade,<br />

1965)............................................................................................................................... 10<br />

Fig. 5 - Syn<strong>the</strong>sis <strong>of</strong> <strong>paraquat</strong>......................................................................................... 14<br />

Fig. 6 - Paraquat dealkylation in alkaline solutions. ...................................................... 15<br />

Fig. 7 – Intermediate resonance structures <strong>of</strong> <strong>paraquat</strong> and full reduction. ................... 16<br />

Fig. 8 - Chemical structure <strong>of</strong> <strong>paraquat</strong> (A) and putrescine (B), showing geometric<br />

standards <strong>of</strong> <strong>the</strong> distance between N atoms (optimal distance to fit polyamine uptake<br />

system is unknown). The chemical structure <strong>of</strong> cadaverin (C), spermidine (D) and<br />

spermine (E) is also presented........................................................................................ 25<br />

Fig. 9 - Autoradiographs <strong>of</strong> rat lung tissue incubated with [ 3 H]putrescine. Resin sections<br />

1 µm thick were stained with toluidine blue and examined by light microscopy.<br />

Labeling occurs in alveolar walls and in alveolar type II pneumocytes (A and B, arrows).<br />

There is no labeling in macrophages (B) or in walls <strong>of</strong> vessels, but Clara cells<br />

xxxi


(arrowheads) in bronchiolar epi<strong>the</strong>lium (C) show intense labeling. Original<br />

magnifications: ×600 in A; ×1,500 in B and C. Adapted from Nemery et al. (Nemery et<br />

al., 1987)......................................................................................................................... 29<br />

Fig. 10 - Autoradiographs <strong>of</strong> human lung tissue incubated with 2.5 µM [ 3 H]putrescine.<br />

Unstained resin sections 1 µm thick were examined by electron spectroscopic imaging.<br />

a-d: 4 different alveolar spaces lined with type II and type I pneumocytes. Silver grains<br />

are evident over type II pneumocytes (long arrows) and lining <strong>of</strong> alveoli (short arrows)<br />

but not over erythrocytes (*) or paranuclear regions <strong>of</strong> endo<strong>the</strong>lium. Cellular and<br />

noncellular components <strong>of</strong> alveolar interstitium were largely devoid <strong>of</strong> silver grains.<br />

Silver grains were uniformly distributed over both nucleus and cytoplasm <strong>of</strong> type II<br />

cells. Bars, 1 µm. Adapted from Hoet et al. (Hoet et al., 1993)..................................... 30<br />

Fig. 11 - Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. A. Cellular<br />

diaphorases, SOD, Superoxide dismutase; CAT, Catalase; GPx, Glutathione Peroxidase;<br />

Gred, Glutathione Reductase; PQ 2+ , Paraquat; PQ •+ , Paraquat monocation free radical;<br />

HMP, Hexose monophosphate pathway, FR; Fenton reaction; HWR, Haber-Weiss<br />

Reaction, PUS; polyamine uptake system...................................................................... 35<br />

Fig. 12 – Pretreatment procedures for <strong>paraquat</strong> in urine and plasma before <strong>the</strong> second<br />

derivative spectrophotometric analysis (A). Zero-order and second-derivative spectrum.<br />

The qualitative analysis is made by observing <strong>the</strong> presence <strong>of</strong> inflection points at about<br />

396 and 403 nm. The quantification is made with amplitudes measurable between 396<br />

and 403 nm (B)............................................................................................................... 63<br />

Fig. 13 - Nomogram showing relation between plasma <strong>paraquat</strong> concentrations<br />

(μg/mL), time after ingestion, and probability <strong>of</strong> survival. Adapted from Hart et al.<br />

(Hart et al., 1984). .......................................................................................................... 65<br />

Fig. 14 – Representative qualitative urinary test for <strong>paraquat</strong>. Correlation between <strong>the</strong><br />

PQ concentration (μg/mL) and <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> blue colour change......................... 68<br />

Fig. 15 – Relationship between urine <strong>paraquat</strong> concentrations and survival. Adapted<br />

from Scherrmann et al. (Scherrmann et al., 1987). ........................................................ 69<br />

xxxii


Fig. 16 - Flowchart guide usually followed for <strong>the</strong> management <strong>of</strong> <strong>paraquat</strong> poisoning.<br />

PQ, <strong>paraquat</strong>; i.v., intravenous; PaO2, partial pressure <strong>of</strong> oxygen in arterial blood; O2,<br />

oxygen; NO, nitric oxide; CHP, charcoal hemoperfusion; CP, cyclophosphamide; MP,<br />

methylprednisolone; DFO, desferoxamine; NAC, N-acetylcysteine; DEX,<br />

dexamethasone; WBC, white-blood-cells; 1 If systemic toxicity is suspected, test urine<br />

for PQ. There is little data for time to peak plasma levels by skin absorption, but if <strong>the</strong><br />

urine is negative for 24 hours, systemic toxicity can probably be disregarded. If <strong>the</strong><br />

urine test is positive or if <strong>the</strong>re is any doubt about potential systemic toxicity, assess<br />

blood concentrations and treat for systemic toxicity as described for ingestion; 2 Risk <strong>of</strong><br />

inducing bleeding, perforation or scarring due to additional trauma to fragilized tissues.<br />

Gastric lavage without administration <strong>of</strong> an adsorbent has not shown any clinical<br />

benefit; 3 Or in 250 mL <strong>of</strong> cathartics (it will increase gut motility to improve excretion <strong>of</strong><br />

<strong>the</strong> charcoal-PQ complex) via nasogastric tube; 4 Maximum dose is 50 g; 5 Repeat doses<br />

<strong>of</strong> cathartics may result in fluid and electrolyte imbalances, particularly in children, and<br />

are <strong>the</strong>refore not recommended; 6 Particularly important as a mean to correct<br />

dehydration, accelerating excretion, reducing tubular concentrations and correcting any<br />

metabolic acidosis. However, fluid balance must be monitored to avoid fluid overload if<br />

renal failure develops. In this case hemodialysis or hem<strong>of</strong>iltration may be required;<br />

7<br />

Plasma should be analyzed ra<strong>the</strong>r than serum, because serum PQ concentrations are<br />

approximately 3 fold lower than those in plasma obtained from <strong>the</strong> same blood sample.<br />

If only serum is available results should be interpreted with caution in relation to<br />

survival curves. Plasma should be stored in plastic and not in glass tubes because PQ 2+<br />

adsorb onto glass surfaces. ............................................................................................. 74<br />

Fig. 17 – Proposed protective <strong>mechanisms</strong> <strong>of</strong> sodium salicylate against pulmonary<br />

<strong>paraquat</strong> toxicity. .......................................................................................................... 215<br />

xxxiii


xxxiv


INDEX OF TABLES<br />

Table 1 – Some <strong>paraquat</strong> trade names. ............................................................................ 4<br />

Table 2 - Countries in which <strong>paraquat</strong> is registered as <strong>of</strong> June 2005. Adapted from<br />

<strong>paraquat</strong> information center (www.<strong>paraquat</strong>.com)........................................................... 5<br />

Table 3 - Physical and chemical properties <strong>of</strong> PQ ion. 1 1 g <strong>of</strong> <strong>paraquat</strong> dichloride =<br />

0.724 g <strong>of</strong> <strong>paraquat</strong> ion................................................................................................... 12<br />

Table 4 - Kinetic constants for <strong>the</strong> accumulation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong> lung tissue slices from<br />

various species Adapted from Rose et al. (Rose et al., 1974)........................................ 23<br />

Table 5 – Phases <strong>of</strong> <strong>paraquat</strong> toxicity and associated clinical effects. GIT,<br />

gastrointestinal tract; CNS, central nervous system; 1 doses as low as 4 mg/Kg can cause<br />

death (Driesbach, 1983).................................................................................................. 41<br />

Table 6 – Paraquat LD50 in various species. NS, not stated; M, male; F, female; a dose<br />

quoted as <strong>paraquat</strong> ion; b as dimethylsulphate................................................................. 47<br />

Table 7 – Clinical features <strong>of</strong> <strong>paraquat</strong> poisonings........................................................ 49<br />

xxxv


xxxvi


PART I<br />

OUTLINE OF THE DISSERTATION<br />

The present <strong>the</strong>sis is structured in four main parts:<br />

1. GENERAL INTRODUCTION<br />

In Part I, <strong>the</strong> present section, a general overview on <strong>the</strong> <strong>research</strong> assumptions,<br />

objectives and structure <strong>of</strong> <strong>the</strong> <strong>the</strong>sis is presented. Considering <strong>the</strong> huge number <strong>of</strong><br />

publications concerning to PQ (more than 4000 referred in “PubMed” database since its<br />

introduction <strong>into</strong> <strong>the</strong> market as an herbicide in 1962), <strong>the</strong> general introduction section<br />

had a focused approach covering aspects <strong>of</strong> <strong>the</strong> clinical features <strong>of</strong> poisoning,<br />

<strong>mechanisms</strong> <strong>of</strong> toxicity, putative treatments and <strong>the</strong>ir relevance to <strong>the</strong> treatment <strong>of</strong><br />

human poisonings. The introduction is restricted principally to those articles in high<br />

standard scientific publications. Considerable space is dedicated to techniques for<br />

prognosis prediction, since <strong>the</strong>se could allow development <strong>of</strong> rigorous clinical protocols<br />

that may produce comparable data for <strong>the</strong> evaluation <strong>of</strong> proposed <strong>the</strong>rapies. It is<br />

expected that this <strong>the</strong>sis may also serve as a guide for clinicians and scientists who work<br />

and study this particularly toxic chemical.<br />

The recent evidences linking PQ exposure with Parkinson’s disease development<br />

are also reviewed in this section. Although <strong>the</strong> original <strong>research</strong> prepared in <strong>the</strong> scope <strong>of</strong><br />

this dissertation was not directed to <strong>the</strong> nervous system, <strong>the</strong> obtained findings may<br />

provide new perspectives and foster fur<strong>the</strong>r investigation on <strong>the</strong> involvement <strong>of</strong> PQ in<br />

Parkinson’s disease.<br />

2. GENERAL AND SPECIFIC OBJECTIVES OF THE DISSERTATION<br />

The general and specific objectives <strong>of</strong> <strong>the</strong> dissertation are provided.<br />

xxxvii


PART II – ORIGINAL RESEARCH<br />

The Part II is divided in six chapters, corresponding to <strong>the</strong> original articles, and<br />

describes <strong>the</strong> experimental work in order to answer <strong>the</strong> questions that derived from <strong>the</strong><br />

general and specific objectives <strong>of</strong> <strong>the</strong> <strong>the</strong>sis.<br />

PART III<br />

This section it is divided in three major points:<br />

1. INTEGRATED OVERVIEW OF THE PERFORMED STUDIES – <strong>the</strong><br />

studies undertaken are integrated in a harmonized form;<br />

2. CONCLUSIONS - <strong>the</strong> conclusions that can be taken from this dissertation are<br />

summarized;<br />

3. DIRECTIONS FOR FUTURE RESEARCH – future studies are projected.<br />

PART IV<br />

The references used in <strong>the</strong> PART I and III are listed.<br />

xxxviii


TABLE OF CONTENTS<br />

AUTHOR’S DECLARATION ....................................................................................VIII<br />

PUBLICATIONS ........................................................................................................... IX<br />

Articles in international peer-reviewed journals ............................................................ IX<br />

Abstracts in international peer-reviewed journals............................................................X<br />

Patents............................................................................................................................. XI<br />

ACKNOWLEDGMENTS............................................................................................XIII<br />

ABSTRACT ..............................................................................................................XVIII<br />

RESUMO ..................................................................................................................XXIII<br />

ABBREVIATIONS LIST ....................................................................................... XXVII<br />

INDEX OF FIGURES.............................................................................................. XXXI<br />

INDEX OF TABLES ..............................................................................................XXXV<br />

OUTLINE OF THE DISSERTATION ................................................................ XXXVII<br />

TABLE OF CONTENTS .......................................................................................XXXIX<br />

PART I<br />

(THEORETICAL BACKGROUND)<br />

1. GENERAL INTRODUCTION ......................................................................... 1<br />

1. HISTORY, USE AND USEFULNESS OF PARAQUAT........................................... 3<br />

1.1 Mode <strong>of</strong> action as herbicide ................................................................................ 7<br />

1.2 Biodegradation Pathways.................................................................................... 8<br />

1.2.1 Photochemical Degradation.................................................................. 9<br />

1.2.1.1 On plant surfaces.............................................................................. 9<br />

1.2.1.2 On soil and o<strong>the</strong>r mineral surfaces ................................................ 10<br />

1.2.2 Microbial degradation ........................................................................ 11<br />

2. CHEMISTRY OF PARAQUAT ................................................................................ 11<br />

2.1 Physical and chemical properties ...................................................................... 12<br />

2.2 Syn<strong>the</strong>sis ........................................................................................................... 13<br />

2.3 Electrochemistry <strong>of</strong> viologens and <strong>paraquat</strong> reduction..................................... 15<br />

3. TOXICOKINETICS OF PARAQUAT...................................................................... 17<br />

3.1 Absorption......................................................................................................... 17<br />

3.2 Distribution ....................................................................................................... 19<br />

3.2.1 Preferential accumulation in <strong>the</strong> lung................................................. 22<br />

3.2.2 Lung accumulation through <strong>the</strong> polyamine uptake system ................ 22<br />

3.2.3 Structural requirements for <strong>the</strong> pulmonary polyamine uptake system26<br />

3.2.4 Characterization <strong>of</strong> <strong>the</strong> pulmonary polyamine uptake system............ 27<br />

3.2.5 Cellular localization <strong>of</strong> <strong>the</strong> polyamine uptake system in <strong>the</strong> lung...... 28<br />

xxxix


3.3 Metabolism........................................................................................................ 31<br />

3.4 Elimination........................................................................................................ 32<br />

4. BIOCHEMICAL MECHANISMS OF PARAQUAT TOXICITY............................ 34<br />

4.1 Mechanism <strong>of</strong> toxicity....................................................................................... 34<br />

4.2 Biochemical consequences <strong>of</strong> <strong>the</strong> redox cycling process ................................. 36<br />

4.2.1 Oxidation <strong>of</strong> NADPH ......................................................................... 36<br />

4.2.2 Oxidation <strong>of</strong> cellular thiol (SH) groups.............................................. 37<br />

4.2.3 Oxidative damage to lipids, proteins and DNA.................................. 38<br />

5. LUNG PATHOPHYSIOLOGY ................................................................................. 40<br />

5.1 Destructive phase .............................................................................................. 42<br />

5.2 Proliferative phase............................................................................................. 44<br />

6. OBSERVATIONS IN ANIMALS AND HUMANS ................................................. 46<br />

6.1 Clinical symptoms and manifestations <strong>of</strong> <strong>paraquat</strong> <strong>into</strong>xication ...................... 49<br />

6.1.1 Poisoning by <strong>the</strong> oral route................................................................. 50<br />

6.1.1.1 Severe toxicity................................................................................. 50<br />

6.1.1.2 Moderate toxicity – <strong>the</strong> typical <strong>into</strong>xication................................... 51<br />

6.1.1.2.1 First phase................................................................................. 51<br />

6.1.1.2.2 Second phase ............................................................................ 52<br />

6.1.1.2.3 Third phase ............................................................................... 53<br />

6.1.1.3 Mild Toxicity................................................................................... 55<br />

6.1.2 Exposure by dermal route................................................................... 55<br />

6.1.3 Ocular irritation .................................................................................. 57<br />

6.1.4 Exposure by inhalation ....................................................................... 58<br />

6.1.5 Muscle toxicity ................................................................................... 58<br />

6.2 Intoxications during pregnancy......................................................................... 58<br />

6.3 Incidents <strong>of</strong> pet animals poisoning.................................................................... 59<br />

7. PREDICTING HUMAN OUTCOME IN PARAQUAT POISONING..................... 59<br />

7.1 Paraquat quantification in biological samples................................................... 60<br />

7.1.1 Qualitative and semi-quantitative test ................................................ 61<br />

7.1.2 Quantitative test: spectrophotometry.................................................. 61<br />

7.1.2.1 Reagents and <strong>the</strong>ir preparation...................................................... 62<br />

7.4.2.2 Analytical conditions...................................................................... 62<br />

7.1.2.3 Procedures...................................................................................... 62<br />

7.2 Predicting <strong>the</strong> outcome from plasma <strong>paraquat</strong> concentrations.......................... 64<br />

7.3 Predicting outcome from urine <strong>paraquat</strong> concentrations................................... 68<br />

7.4 Additional laboratory tests ................................................................................ 69<br />

8. TREATMENT............................................................................................................ 72<br />

8.1 Preventing <strong>paraquat</strong> absorption......................................................................... 72<br />

8.2 Increasing <strong>paraquat</strong> elimination ........................................................................ 76<br />

8.2.1 Extracorporeal elimination ................................................................. 76<br />

8.2.2 Forced diuresis and peritoneal dialysis............................................... 78<br />

8.3 Supportive <strong>the</strong>rapies.......................................................................................... 79<br />

8.4 Measures to prevent lung damage..................................................................... 80<br />

8.5 New perspectives .............................................................................................. 94<br />

8.5.1 Mechanical ventilation with additional inhalation <strong>of</strong> NO .................. 95<br />

8.5.2 Prop<strong>of</strong>ol .............................................................................................. 96<br />

9. SEQUELAE IN SURVIVORS................................................................................... 96<br />

10. LUNG APPEARANCE AT AUTOPSY .................................................................. 96<br />

11. REVIEW ARTICLE - PARAQUAT EXPOSURE AS AN ETIOLOGICAL<br />

FACTOR OF PARKINSON'S DISEASE...................................................................... 99<br />

xl


2. GENERAL AND SPECIFIC OBJECTIVES OF THE DISSERTATION ...... 115<br />

PART II<br />

(ORIGINAL RESEARCH)<br />

CHAPTER I.................................................................................................... 123<br />

CHAPTER II................................................................................................... 141<br />

CHAPTER III.................................................................................................. 147<br />

CHAPTER IV.................................................................................................. 161<br />

CHAPTER V................................................................................................... 177<br />

CHAPTER VI.................................................................................................. 191<br />

PART III<br />

1. INTEGRATED OVERVIEW OF THE PERFORMED STUDIES ................. 207<br />

2. CONCLUSIONS......................................................................................... 217<br />

3. DIRECTIONS FOR FUTURE RESEARCH ................................................ 221<br />

PART IV<br />

1. REFERENCES........................................................................................... 225<br />

xli


xlii


1. GENERAL INTRODUCTION<br />

PART I<br />

1. GENERAL INTRODUCTION<br />

1


Part I - General Introduction__________________________________________________<br />

2


__________________________________________________Part I - General Introduction<br />

1. HISTORY, USE AND USEFULNESS OF PARAQUAT<br />

Paraquat (1,1′-dimethyl-4,4′-bipyridylium dichloride; PQ) [CAS #1910-42-5] is an<br />

herbicide belonging to <strong>the</strong> chemical family <strong>of</strong> bipyridylium (also called bipyridyl)<br />

quaternary ammonium herbicides. PQ and diquat (1,1′-ethylene-2,2′-dipyridylium<br />

dibromide; DQ) [CAS #85-00-7] are <strong>the</strong> most commonly used herbicides <strong>of</strong> this group<br />

(Fig. 1). They have similar chemical and physical properties and have a similar mode <strong>of</strong><br />

action on plants (Calderbank, 1968). Of <strong>the</strong> two bipyridylium herbicides (BHs) in use,<br />

PQ is by far <strong>the</strong> most clinically significant in terms <strong>of</strong> number <strong>of</strong> <strong>into</strong>xication cases, and<br />

it will be <strong>the</strong> main subject <strong>of</strong> this introduction.<br />

H 3<br />

C<br />

+ +<br />

N<br />

N CH 3<br />

1.02 nm<br />

PQ (1,1´-dimethyl-4,4´-bipyridylium ion)<br />

+ +<br />

N N<br />

DQ (1,1′-ethylene-2,2′-dipyridylium ion)<br />

Fig. 1 – Chemical structures <strong>of</strong> <strong>paraquat</strong> (PQ) and diquat (DQ).<br />

PQ was first described in 1882 (Weidel and Rosso, 1882). Its redox properties<br />

were discovered in 1933 by Michaelis and Hill (Michaelis and Hill, 1933). By that time<br />

it was used as an oxidation-reduction indicator because an electron donation to <strong>the</strong> PQ<br />

ion (PQ 2+ ) forms a stable free radical monocation (PQ •+ ) having a violet or blue colour<br />

(Michaelis and Hill, 1933); hence, PQ is commonly called methyl viologen (Fig. 1).<br />

3


Part I - General Introduction __________________________________________________<br />

4<br />

The PQ herbicidal properties were discovered at <strong>the</strong> Jealott’s Hill International<br />

Research Centre, Bracknell, UK in 1955, and in August 1962, PQ was introduced <strong>into</strong><br />

<strong>the</strong> market as an herbicide by <strong>the</strong> Plant Protection Division Ltd <strong>of</strong> Imperial Chemical<br />

Industries [(ICI), now Syngenta] (Homer et al., 1960; Calderbank, 1968; Smith and<br />

Heath, 1976). Gramoxone®, manufactured by Syngenta, is <strong>the</strong> most common trade<br />

name for PQ, but <strong>the</strong> herbicide is also sold under many different trade names by several<br />

different companies (Table 1).<br />

Table 1 – Some <strong>paraquat</strong> trade names.<br />

Paraquat Paraquat-Diquat Mixtures<br />

Crisquat Preeglone<br />

Dextrone X Priglone<br />

Esgram Weedol<br />

Gramoxone<br />

In spite <strong>of</strong> <strong>the</strong> numerous <strong>into</strong>xications, PQ is now registered and used in over 120<br />

developed and developing countries throughout <strong>the</strong> world (Table 2). The main reasons<br />

for such widespread use are <strong>the</strong> following:<br />

• PQ is an excellent herbicide for destroying weeds that may decrease crop<br />

yields, during normal application. It is also used in pasture renovation and on<br />

non-crop areas such as public airports, electronic transformer stations, and<br />

around commercial buildings. Its success as a weed-killer lies on <strong>the</strong> fact that<br />

small quantities <strong>of</strong> a PQ solution will rapidly kill plants on contact with <strong>the</strong><br />

leaves and due to its low cost. In addition, PQ allows <strong>the</strong> roots to remain intact,<br />

thus holding <strong>the</strong> soil toge<strong>the</strong>r and preventing soil erosion;<br />

• PQ is highly hydrophilic and thus not absorbed through intact skin;<br />

• Aerosolized PQ particles are larger than 5 μm in diameter and thus do not reach<br />

<strong>the</strong> humans alveoli when exposed by inhalation route;<br />

• PQ is rapidly inactivated and metabolized once in <strong>the</strong> soil, preventing its<br />

accumulation in <strong>the</strong> ecosphere.


__________________________________________________Part I - General Introduction<br />

Table 2 - Countries in which <strong>paraquat</strong> is registered as <strong>of</strong> June 2005. Adapted from<br />

<strong>paraquat</strong> information center (www.<strong>paraquat</strong>.com).<br />

Albania El Salvador Malaysia Sierra Leone<br />

Algeria Ethiopia Mali Singapore<br />

Angola Fiji Malta Slovakia<br />

Antigua & Barbuda France Mauritania Somalia<br />

Argentina Gabon Mauritius South Africa<br />

Australia Gambia Mexico South Korea<br />

Bahamas Germany Morocco Spain<br />

Bahrain Ghana Mozambique Sri Lanka<br />

Bangladesh Greece Myanmar Sudan<br />

Barbados Grenada Namibia Suriname<br />

Belgium Guatemala Ne<strong>the</strong>rlands Swaziland<br />

Belize Guinea New Zealand Tahiti<br />

Bolivia Guinea-Bissau Nicaragua Taiwan<br />

Botswana Guyana Niger Tanzania<br />

Brazil Haiti Nigeria Thailand<br />

Burkina Faso Honduras Oman Trinidad & Tobago<br />

Burundi India Pakistan Turkey<br />

Cameroon Indonesia Panama Uganda<br />

Canada Iran Papua New Guinea United Kingdom<br />

Cape Verde Iraq Paraguay USA<br />

Chad Ireland Peru Uruguay<br />

Chile Israel Philippines Venezuela<br />

China Italy Poland Vietnam<br />

Colombia Jamaica Portugal Yemen<br />

Costa Rica Japan Romania Yugoslavia<br />

Cote d’Ivoire Jordan Rwanda Zambia<br />

Croatia Kenya Sao Tome & Principe Zimbabwe<br />

Cuba Lebanon St Kitts & Nevis<br />

Czech Republic Liberia St Lucia<br />

Dominica Macedonia St Vincent & Grenadines<br />

Dominican Republic Madagascar Samoa<br />

Ecuador Malawi Senegal<br />

Since its introduction in <strong>the</strong> market, numerous successful practical uses <strong>of</strong> <strong>the</strong><br />

herbicide have been implemented. PQ is an extremely effective, fast-acting and non-<br />

selective foliage-applied contact herbicide, killing a wide range <strong>of</strong> grass and dicot<br />

weeds. It is a defoliant, desiccant, and plant growth regulator herbicide. PQ is rapidly<br />

inactivated by <strong>the</strong> majority <strong>of</strong> surrounding soils in <strong>the</strong> event <strong>of</strong> overspray (Amondham<br />

et al., 2006). Inactivation on contact with soil means that no biologically active residues<br />

remain in <strong>the</strong> soil, thus allowing planting or sowing to be carried out almost<br />

immediately after spraying. The PQ 2+ is strongly attracted to <strong>the</strong> negative charge <strong>of</strong> soil<br />

clay particles and once <strong>the</strong> equilibrium is established (Fig. 2), PQ, at typical<br />

5


Part I - General Introduction __________________________________________________<br />

environmentally expected concentrations, becomes a strongly adsorbed residue that is<br />

biologically unavailable due to having an extremely low concentration in <strong>the</strong> soil<br />

solution. The soil’s natural deactivation capacity for this herbicide is several times <strong>the</strong><br />

normally recommended application rate (Smith and Oehme, 1991), existing evidences<br />

demonstrating that adsorption is capable <strong>of</strong> deactivating <strong>the</strong> equivalent <strong>of</strong> hundreds or<br />

even thousands <strong>of</strong> PQ applications over a wide range <strong>of</strong> soils. This also means that PQ<br />

is effectively immobilized in soils with no leaching to ground water (Roberts et al.,<br />

2002).<br />

6<br />

Soil Surfaces Soil Pore Water<br />

>99.99% Microbial<br />

Strongly Adsorbed<br />

Degradation<br />

CO 2 + H 2 O<br />

Fig. 2 – Equilibrium dynamics for <strong>paraquat</strong> between <strong>the</strong> soil and soil solution. Adapted<br />

from Roberts et al. (Roberts et al., 2002).<br />

Although <strong>the</strong> non-systemic (contact) property <strong>of</strong> PQ makes it less than ideal for<br />

<strong>the</strong> long-term control <strong>of</strong> perennial weeds, <strong>the</strong> same property is <strong>of</strong> real advantage when<br />

parts <strong>of</strong> crop plants are sprayed accidentally and thus only <strong>the</strong> part receiving <strong>the</strong> spray is<br />

affected (Sagar, 1987). PQ is only rainfast within minutes <strong>of</strong> application. This property<br />

reduces <strong>the</strong> operator’s dependence on wea<strong>the</strong>r and allows great precision in <strong>the</strong> timing<br />

<strong>of</strong> applications. In relation to <strong>the</strong> crop, PQ may be applied preharvest, preemergence, or<br />

preplant. The herbicidal activity becomes obvious as a rapid decolourization and<br />

desiccation <strong>of</strong> green plant tissue when illuminated. PQ has also been used for control <strong>of</strong><br />

aquatic weeds in irrigation ditches from where residues disappear rapidly due to its<br />

strong adsorption to bottom mud and onto aquatic weeds (Grover et al., 1980). The best<br />

herbicidal results are achieved by spraying in late afternoon or evening. However, <strong>the</strong><br />

herbicidal activity is slower in <strong>the</strong> dark, owing to <strong>the</strong> absence <strong>of</strong> naturally occurring<br />

reducing agents during photosyn<strong>the</strong>sis.<br />

Product formulations differ among countries. Typically, it is available as a 10% to<br />

30% concentrated solution (according to <strong>the</strong> manufacturer’s instructions, correctly<br />

diluted spray solutions should contain no more than 0.05 to 0.2% <strong>of</strong> PQ ion) for<br />

agricultural use or as a 2.5 or 5% powder (w/w) for domestic use. Granular and gel<br />

forms are also encountered. PQ is caustic, and <strong>the</strong> concentrate may also contain an


__________________________________________________Part I - General Introduction<br />

aliphatic detergent to enhance entry <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> cells and thus its toxicity. PQ can be<br />

applied safely when used according to <strong>the</strong> manufacturer’s guidelines (Hart, 1987). It is<br />

generally with <strong>the</strong> liquid formulations (specially <strong>the</strong> concentrated ones) for agricultural<br />

use that <strong>the</strong> vast majority <strong>of</strong> fatal cases <strong>of</strong> <strong>into</strong>xication have occurred. Proudfoot et al.<br />

(Proudfoot et al., 1987) reported a mortality <strong>of</strong> 65% in patients who ingested <strong>the</strong><br />

concentrated formulation and only 4% in those who ingested <strong>the</strong> diluted solutions (2.5%<br />

w/v). In <strong>the</strong> granular form, it is difficult to ingest by accident, and large quantities <strong>of</strong> <strong>the</strong><br />

granules have to be ingested to induce toxic effects. Originally, marketed aqueous PQ<br />

formulations were brown in colour. Due to mistakes with o<strong>the</strong>r common beverages such<br />

as c<strong>of</strong>fee, cola drinks, among o<strong>the</strong>rs, <strong>the</strong> colour is now dark blue-green. The<br />

formulations also contain a powerful stenching and emetic agent. Recently, Syngenta<br />

scientists have developed a formulation, Gramoxone Inteon®, which contains a gelling<br />

agent (alginate) that is activated or triggered at <strong>the</strong> pH <strong>of</strong> stomach acid, increased levels<br />

<strong>of</strong> emetic and a purgative. Once formed, <strong>the</strong> gel minimizes and slows dispersion and<br />

passage <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> small intestines, thus allowing more time for productive emesis<br />

as caused by <strong>the</strong> emetic. The new formulation improved overall survival following PQ<br />

ingestion from 25.6% to 35.3% (www.<strong>paraquat</strong>.com).<br />

1.1 Mode <strong>of</strong> action as herbicide<br />

Herbicidal activity as well PQ-<strong>induced</strong> toxicity to mammals was found to be<br />

linked to PQ redox potential (Bird and Kuhn, 1981). Initial work on <strong>the</strong> mode <strong>of</strong> action<br />

<strong>of</strong> BHs by Mees (Mees, 1960) indicated that <strong>the</strong>ir ability to cause rapid kill is dependent<br />

on <strong>the</strong> photosyn<strong>the</strong>tic activity <strong>of</strong> plants and on oxygen (O2). Zweig et al. (Zweig et al.,<br />

1965) fur<strong>the</strong>r found that BHs cause a deviation <strong>of</strong> electron flow from Photosystem I<br />

(which normally transfers its electron to ferredoxin), leading to an inhibition <strong>of</strong> oxidized<br />

nicotinamide adenine dinucleotide phosphate (NADP + ) reduction during photosyn<strong>the</strong>sis<br />

(Fig. 3). As result <strong>of</strong> this process, a PQ •+ is produced in <strong>the</strong> cell at <strong>the</strong> expense <strong>of</strong><br />

NADPH. Thus, PQ is only toxic to <strong>the</strong> green parts <strong>of</strong> <strong>the</strong> plant, where <strong>the</strong><br />

photosyn<strong>the</strong>sis occurs (Slade, 1966). When plants are irradiated with sunlight, <strong>the</strong><br />

generated electrons reduce PQ 2+ to PQ •+ , which is rapidly reoxidized by <strong>the</strong> O2<br />

produced in chloroplasts (Slade, 1966). During <strong>the</strong> reoxidization a superoxide radical<br />

(O2 .- ) is generated, with <strong>the</strong> subsequent deleterious effects and consequent cell death.<br />

7


Part I - General Introduction __________________________________________________<br />

Therefore <strong>the</strong> mechanism <strong>of</strong> toxic action <strong>of</strong> PQ involves cyclic reduction–oxidation<br />

reactions, which produce reactive oxygen species (ROS) and depletion <strong>of</strong> reduced<br />

nicotinamide adenine dinucleotide phosphate (NADPH). An excellent review <strong>of</strong> <strong>the</strong><br />

subject is given by Dodge (Dodge, 1971).<br />

Fig. 3 – Herbicidal mechanism <strong>of</strong> <strong>paraquat</strong>. In photosystem I (PS I), plastocyanin (PC)<br />

transfers its electron (e - ) through a series <strong>of</strong> steps (P700, A0, A1, FX, FA) to ferrodoxin<br />

(FRD) and finally to NADP + . PQ ion (PQ 2+ ) binds near <strong>the</strong> FRD binding site in PS I and<br />

accepts an e - , becoming <strong>paraquat</strong> monocation free radical (PQ •+ ), which initiates a series<br />

<strong>of</strong> reactions leading to cell membrane disruption and plant death. The formation <strong>of</strong> such<br />

free radicals stops electron transport to NADP + and effectively inhibits normal<br />

functioning <strong>of</strong> PS I. A, Ferrodoxin-NADP + reductase.<br />

1.2 Biodegradation Pathways<br />

When bound to soil, PQ is biologically inactive, unavailable for ei<strong>the</strong>r herbicidal<br />

or ecotoxicological action and unavailable for microbial degradation or<br />

8<br />

Stroma<br />

Lumen<br />

PQ 2+<br />

PsaD<br />

hv<br />

PsaA<br />

FA FB<br />

e- e- e- e- e- e- e- e- PC 2+<br />

F X<br />

A 1<br />

A 0<br />

P 700<br />

PC +<br />

NADP + + H +<br />

A<br />

X<br />

FRD ox FRD red<br />

e- e- PsaB<br />

PsaF<br />

PsaE<br />

PsaC<br />

PQ .+<br />

PSI<br />

NADPH<br />

O 2<br />

O 2 .-<br />

PQ 2+<br />

CELLULAR<br />

DAMAGE


__________________________________________________Part I - General Introduction<br />

photodecomposition (Burns and Audus, 1970; Smith and Oehme, 1991). The strong<br />

binding <strong>of</strong> PQ to clay minerals (e.g., bentonite) forms <strong>the</strong> basis <strong>of</strong> a suggested method<br />

for preventing systemic absorption in human poisoning cases (Smith et al., 1974a).<br />

Such soil-bound residues may persist essentially indefinitely, with only a 10% annual<br />

loss and a field half-life (t1/2) <strong>of</strong> 6.6 years (Hance et al., 1980). PQ is only significantly<br />

available for degradation during <strong>the</strong> immediate period after soil application (especially<br />

during <strong>the</strong> first 96 hours), when <strong>the</strong> herbicide is only weakly adsorbed to <strong>the</strong> soil<br />

particles (Burns and Audus, 1970).<br />

1.2.1 Photochemical Degradation<br />

The photochemical degradation <strong>of</strong> PQ has been observed in laboratory conditions,<br />

as well as on <strong>the</strong> surface <strong>of</strong> plant matter and soils. It is <strong>the</strong> predominant mechanism <strong>of</strong><br />

PQ degradation in soils (Smith and Mayfield, 1978) and it is related to <strong>the</strong> availability<br />

<strong>of</strong> UV between <strong>the</strong> wavelengths <strong>of</strong> 290 and 310 nm during daylight hours (Slade, 1965;<br />

Slade, 1966). The main intermediates <strong>of</strong> photochemical PQ degradation on plants or soil<br />

surfaces are <strong>of</strong> low toxicity. They decompose easily and are not expected to produce<br />

adverse environmental effects.<br />

1.2.1.1 On plant surfaces<br />

In agricultural practice, much <strong>of</strong> <strong>the</strong> sprayed PQ is initially deposited on plant<br />

surfaces. Slade (Slade, 1965; Slade, 1966) applied PQ dichloride droplets to maize,<br />

tomato, and broad-bean plants and studied <strong>the</strong> degradation pathways. Determinations<br />

carried out at intervals <strong>of</strong> 100 days showed that degradation was caused by<br />

photochemical decomposition on <strong>the</strong> leaf surfaces and not by metabolism. Degradation<br />

products isolated from plants sprayed with [ 14 C]PQ dichloride included 4-carboxyl-1methyl-<br />

14 C-pyridylium chloride or 4-methylisonicotinic acid (MINA) and methylamine-<br />

14<br />

C-hydrochloride. The photochemical degradation <strong>of</strong> PQ dichloride continued after <strong>the</strong><br />

plants were dead (Fig. 4). The photochemical degradation <strong>of</strong> PQ is rapid. A 0.1% PQ<br />

solution was completely degraded in 3 days under a UV lamp (Slade, 1965). PQ<br />

photodegradation products were not translocated from <strong>the</strong> desiccated leaves <strong>of</strong> <strong>the</strong><br />

9


H3C<br />

O<br />

N<br />

Part I - General Introduction __________________________________________________<br />

plants, nor were <strong>the</strong>y found in <strong>the</strong> crops (cereals and fruits), when weeds were treated<br />

with PQ during 3-4 successive seasons (Slade, 1965).<br />

10<br />

O<br />

N<br />

CH 3<br />

C<br />

H 3<br />

C<br />

H 3<br />

O 2 , UV radiation<br />

+ +<br />

N<br />

N CH3 + +<br />

N<br />

N<br />

C N<br />

PQ monopyridone<br />

O<br />

O 2 , UV radiation<br />

C<br />

H 3<br />

CH 3<br />

+<br />

N<br />

microbial<br />

H 3<br />

NH 3 + CO 2 + H 2 O<br />

O 2 , UV radiation, microbial<br />

COOH<br />

N-methylisonicotinic acid (MINA)<br />

Monoquat<br />

O 2 , UV radiation, microbial<br />

CO 2 + CH 3 NH 2 HCl + formate + oxalate + succinate<br />

O 2 , UV radiation<br />

Fig. 4 – Photochemical and microbial degradation <strong>of</strong> PQ. Adapted from Slade (Slade,<br />

1965).<br />

1.2.1.2 On soil and o<strong>the</strong>r mineral surfaces<br />

Slade (Slade, 1966) showed that <strong>the</strong>re was a breakdown, similar to that observed<br />

on plant surfaces, if spots <strong>of</strong> PQ on silica gel were directly exposed to sunlight. When<br />

[ 14 C]PQ dichloride was sprayed on <strong>the</strong> bare soil surface <strong>of</strong> a field during a hot sunny<br />

period, traces <strong>of</strong> MINA were detected in <strong>the</strong> top inch <strong>of</strong> soil for <strong>the</strong> first few weeks<br />

afterwards (Calderbank and Slade, 1976). Radioassay showed that <strong>the</strong> total soil residue<br />

did not markedly decrease during a 6-18 month period, so that, in agricultural practice,<br />

UV degradation <strong>of</strong> herbicide reaching <strong>the</strong> soil should be regarded as insignificant.<br />

N<br />

CO 2 + CH 3 NH 2 HCl


__________________________________________________Part I - General Introduction<br />

1.2.2 Microbial degradation<br />

Microbial PQ degradation has been thoroughly reviewed by Haley (Haley, 1979).<br />

In soils, it does not occur at appreciable rates due to <strong>the</strong> sequestering <strong>of</strong> <strong>the</strong> herbicide at<br />

mineral or <strong>organ</strong>ic anionic sites. Never<strong>the</strong>less, <strong>the</strong> biodegradation <strong>of</strong> PQ has been<br />

observed by a wide variety <strong>of</strong> soil micro<strong>organ</strong>isms in aqueous solution. Baldwin et al.<br />

(Baldwin et al., 1966) identified many soil micro<strong>organ</strong>isms capable <strong>of</strong> degrading PQ.<br />

The herbicide was decomposed by Corynebacterium fascians, Clostridium<br />

pasteurianum, and Lipomyces starkeyi. Several o<strong>the</strong>r micro<strong>organ</strong>isms were found to<br />

degrade PQ (Smith and Heath, 1976) but Lipomyces starkeyi proved to be <strong>the</strong> most<br />

active (Burns and Audus, 1970). In pure culture, degradation <strong>of</strong> PQ has been reported to<br />

occur under both aerobic and anaerobic conditions by Clostridium pasteurianum<br />

(Baldwin et al., 1966). In <strong>the</strong> case <strong>of</strong> <strong>the</strong> yeast Lipomyces starkeyi, <strong>the</strong> ability to degrade<br />

PQ was only seen under aerobic conditions (Funderburk, 1969).<br />

Several decomposition products <strong>of</strong> PQ have been characterized. The demethylated<br />

product, 1-methyl,-4,4’-bipyridylium (Fig. 4) was recovered from an unidentified<br />

bacterial culture as well as <strong>the</strong> N-methyl betaine <strong>of</strong> isonicotinic acid or MINA<br />

(Summers, 1980). Whe<strong>the</strong>r <strong>the</strong>se compounds occur sequentially in one degradation<br />

pathway or in separate pathways is unknown. The N-methyl betaine has been shown to<br />

be readily degraded in soils <strong>into</strong> methylamine and CO2 by microbial activity (Wright<br />

and Cain, 1970). The pyridylium ring carbons are known to be lost as CO2 by 14 C<br />

labelling studies. The methylamine can be utilized as a source <strong>of</strong> nitrogen and carbon.<br />

However, <strong>the</strong> enzymology <strong>of</strong> <strong>the</strong> degradation <strong>of</strong> PQ has not been reported and <strong>the</strong> o<strong>the</strong>r<br />

intermediates have not been identified.<br />

2. CHEMISTRY OF PARAQUAT<br />

The chemistry <strong>of</strong> PQ is dominated by its ability to act as a one-electron carrier.<br />

The electron can be transferred to PQ ei<strong>the</strong>r partially (from a nucleophile or o<strong>the</strong>r<br />

electron-rich compound) or completely (from a reducing agent). In <strong>the</strong> first case,<br />

coloured charge-transfer complexes are formed; in <strong>the</strong> second case, <strong>the</strong> blue-coloured<br />

PQ •+ is formed.<br />

11


Part I - General Introduction __________________________________________________<br />

2.1 Physical and chemical properties<br />

12<br />

The physical and chemical properties <strong>of</strong> PQ 2+ are summarized in Table 3. PQ is<br />

highly water soluble, slightly soluble in alcohol and practically insoluble in <strong>organ</strong>ic<br />

solvents (Haley, 1979). PQ is non-explosive and non-flammable in aqueous<br />

formulations. It is corrosive to metals and incompatible with alkylarylsulfonate wetting<br />

agents and strong oxidizing substances. Although non-ionic surfactants may be used in<br />

combination, PQ is inactivated by anionic ones. It is stable in acid or neutral solutions<br />

but is readily hydrolysed by alkaline solutions (at pH >12). In <strong>the</strong> original container and<br />

under normal conditions, <strong>the</strong> shelf life <strong>of</strong> PQ is indefinitely long and it is also stable at<br />

temperatures above <strong>the</strong> general environmental range.<br />

Table 3 - Physical and chemical properties <strong>of</strong> PQ ion. 1 1 g <strong>of</strong> <strong>paraquat</strong> dichloride =<br />

0.724 g <strong>of</strong> <strong>paraquat</strong> ion.<br />

Class bipyridylium herbicide<br />

Molecular formula C12H14N2<br />

Molecular weight 186.3 (ion), 257.2 (dichloride) 1<br />

Common name <strong>paraquat</strong><br />

IUPAC name 1,1-dimethyl-4,4-bipyridilium<br />

CAS name 1,1-dimethyl-4,4-bipyridilium<br />

Synonyms methyl viologen<br />

CAS Nº<br />

4685-14-7 (ion), 1910-42-5 (dichloride) and<br />

2074-50-2 (sulphate)<br />

Specific gravity (20ºC) 1.240-1.260 g/cm 3<br />

Physical state<br />

Melting point<br />

Boiling point<br />

Solubility in water at 20°C 700 g/L<br />

pH <strong>of</strong> liquid formulation 6.5-7-5<br />

white (pure salts), yellow (technical products)<br />

crystalline, odorless, hygroscopic powders<br />

PQ dichloride melts with decomposition at<br />

~340ºC to form poisonous vapors<br />

PQ dichloride decomposes at ~340ºC to form<br />

poisonous vapors


__________________________________________________Part I - General Introduction<br />

Vapor pressure not measurable<br />

E0’ (relative to <strong>the</strong> normal<br />

hydrogen electrode<br />

Octanol/water partition<br />

coefficient as log Pow (20ºC)<br />

– 0.446 V<br />

-4.2<br />

Dissociation constant PQ ion does not dissociate<br />

Relative gas density 8.88<br />

UV spectrum (in water)<br />

X-ray analysis <strong>of</strong> <strong>the</strong> crystalline<br />

dichloride<br />

Single band centered at 257 nm (Kosower and<br />

Cotter, 1964)<br />

Show two coplanar pyridine rings with two<br />

methyl groups (Russell and Wallwork, 1972)<br />

The basic chemical nucleus <strong>of</strong> PQ (Fig. 1) is a bipyridylium consisting <strong>of</strong> two<br />

quaternized pyridine rings bonded toge<strong>the</strong>r such that <strong>the</strong>ir nitrogen atoms face<br />

diametrically away from one ano<strong>the</strong>r. The quaternization is <strong>the</strong> result <strong>of</strong> <strong>the</strong> methyl<br />

radical addition (para position) to each <strong>of</strong> two nitrogen nuclei in <strong>the</strong> pyridine rings. The<br />

compound is, <strong>the</strong>refore, a para substituted quaternary bipyridylium, hence its common<br />

designation PQ. In its usual oxidized form, it is ionized (bearing two positive charges)<br />

and it is most commonly manufactured as a dichloride salt. Chemically, PQ is thus 1,l’dimethyl-4,4’<br />

dipyridylium dichloride. The positive charges are largely resident on <strong>the</strong><br />

nitrogen atoms as shown by nuclear magnetic resonance (NMR) (Smith and Schneider,<br />

1961). The herbicidal efficacy <strong>of</strong> PQ is related to <strong>the</strong> concentration <strong>of</strong> free PQ 2+ in<br />

solution inside <strong>the</strong> chloroplast, but this will depend (in some cases, markedly) on <strong>the</strong><br />

nature <strong>of</strong> <strong>the</strong> anion and any complexing agent with which it is applied (Homer and<br />

Tomlinson, 1959). For instance, many phenols form soluble crystalline complexes with<br />

PQ dichloride (White, 1969; Ledwith and Woods, 1970). This ready formation <strong>of</strong><br />

complexes is possibly one <strong>of</strong> <strong>the</strong> factors that change <strong>the</strong> herbicidal activity <strong>of</strong> PQ among<br />

floral species and through <strong>the</strong> plant lifespan. Many plants constituents, such lignin and<br />

tannin, are phenolic in nature and could cause immobilization <strong>of</strong> PQ (White, 1969;<br />

Ledwith and Woods, 1970).<br />

2.2 Syn<strong>the</strong>sis<br />

PQ does not occur naturally. It was originally syn<strong>the</strong>sized by Weidel and Rosso as<br />

reported in 1882 (Weidel and Rosso, 1882). There are several methods available for <strong>the</strong><br />

13


Part I - General Introduction __________________________________________________<br />

syn<strong>the</strong>sis <strong>of</strong> PQ. In <strong>the</strong> most common method, PQ is produced by coupling pyridine in<br />

<strong>the</strong> presence <strong>of</strong> sodium in anhydrous ammonia and quaternizing <strong>the</strong> 4,4’-bipyridyl with<br />

an excess <strong>of</strong> methyl chloride to obtain PQ dichloride (Fig. 5). When bipyridyl is<br />

refluxed with methyl iodide or methyl bromide, <strong>the</strong> iodide and <strong>the</strong> bromide salt is<br />

obtained, respectively. The methyl sulfate salt can be obtained by heating 4,4’-bipyridyl<br />

with sodium acetate at 70ºC for 2 hours, <strong>the</strong>n adding methyl sulfate and stirring for 15<br />

min. Haley and Summers (Haley, 1979; Summers, 1980) thoroughly reviewed <strong>the</strong><br />

published methods for PQ syn<strong>the</strong>sis, and for <strong>the</strong> separation and purification <strong>of</strong><br />

bipyridylium salts. The yields obtainable vary from 20% to 96% <strong>of</strong> pure product. The<br />

only impurity permitted in <strong>the</strong> final product is <strong>the</strong> 4,4’-bipyridyl at a maximum level <strong>of</strong><br />

0.25% <strong>of</strong> <strong>the</strong> PQ content (Summers, 1980).<br />

14<br />

2Cl -<br />

+<br />

2 N<br />

N<br />

C<br />

H 3<br />

Fig. 5 - Syn<strong>the</strong>sis <strong>of</strong> <strong>paraquat</strong>.<br />

+<br />

N<br />

+<br />

+<br />

2CH 3 Cl<br />

Na/NH 3 /O 2<br />

N<br />

+<br />

N CH3


C<br />

H 3<br />

__________________________________________________Part I - General Introduction<br />

2.3 Electrochemistry <strong>of</strong> viologens and <strong>paraquat</strong> reduction<br />

The viologens exist in three main oxidation states, namely V 2+ ↔ V •+ → V. The<br />

first reduction step is highly reversible and can be cycled many times without<br />

significant side reaction. The fur<strong>the</strong>r reduction to <strong>the</strong> fully reduced state is less<br />

reversible, not only because <strong>the</strong> latter is frequently insoluble but also because it is an<br />

uncharged one. The compounds are also very stable chemically, although in more<br />

alkaline solutions <strong>the</strong>y will dealkylate (Fig. 6) as reported by Farrington et al.<br />

(Farrington et al., 1969). Because <strong>the</strong> methanol resulting from <strong>the</strong> dealkylation can be a<br />

reducing agent, solutions <strong>of</strong> methyl viologen in alkali can spontaneously be reduced and<br />

will <strong>the</strong>n turn blue as <strong>the</strong> PQ •+ is formed. Methanol is <strong>the</strong>n itself oxidized to<br />

formaldehyde.<br />

OH - CH 3 OH<br />

+ +<br />

+<br />

N<br />

N CH N<br />

N CH 3<br />

3<br />

Fig. 6 - Paraquat dealkylation in alkaline solutions.<br />

The PQ divalent cation is colorless, whereas <strong>the</strong> partially reduced PQ •+ is blue<br />

colored and contains an odd electron. The odd electron is shared by all <strong>the</strong> nuclear<br />

carbon positions in <strong>the</strong> rings (Calderbank, 1968). This step is completely reversible,<br />

such that one equivalent <strong>of</strong> a reducing agent will reduce more than 50% <strong>of</strong> PQ 2+ to PQ •+<br />

only if its reduction potential is more negative than that <strong>of</strong> PQ. Ito and Kuwana (Ito and<br />

Kuwana, 1971) quoted <strong>the</strong> potential <strong>of</strong> <strong>the</strong> first reduction for PQ 2+ as -0.446 V and <strong>the</strong><br />

second is given as -0.88 V (relative to normal hydrogen electrode). A suitable reducing<br />

agent for generating <strong>the</strong> radical is sodium dithionite in alkaline solution (-1.13 V).<br />

Considering that PQ •+ carries an unpaired electron in a π anti-bonding orbital, PQ •+ is<br />

remarkably unreactive. In addition, unpaired electron in PQ •+ is not fixed.<br />

Delocalization <strong>of</strong> <strong>the</strong> unpaired electron gives considerable resonance stabilization to <strong>the</strong><br />

radical and thus it may diffuse outside <strong>the</strong> cell before reacting with O2 (Fig. 7). It is a<br />

known fact that <strong>the</strong> greater <strong>the</strong> number <strong>of</strong> positive centres available to an odd electron,<br />

<strong>the</strong> greater <strong>the</strong> number <strong>of</strong> resonance structures and <strong>the</strong> higher <strong>the</strong> stability <strong>of</strong> <strong>the</strong> free<br />

15


Part I - General Introduction __________________________________________________<br />

radical. For review see (Akhavein and Linscott, 1968). The second step <strong>of</strong> PQ reduction<br />

is <strong>the</strong> addition <strong>of</strong> a second electron to <strong>the</strong> molecule to originate <strong>the</strong> 1,l’-dimethyl-4,4’dihydrobipyridyl.<br />

If it is required to stop <strong>the</strong> reduction at <strong>the</strong> radical stage, ei<strong>the</strong>r a<br />

limited amount <strong>of</strong> <strong>the</strong> reducing agent must be used or <strong>the</strong> solution must be “poised” to<br />

<strong>the</strong> desired reducing potential (-0.56 V for 99% conversion to radical) by adjusting <strong>the</strong><br />

concentration <strong>of</strong> <strong>the</strong> reductant or <strong>the</strong> pH if relevant. The resulting fully reduced species<br />

is colourless (Michaelis and Hill, 1933).<br />

16<br />

C<br />

H 3<br />

C<br />

H 3<br />

C<br />

H 3<br />

C<br />

H 3<br />

+<br />

N<br />

N<br />

+ +<br />

N<br />

N<br />

.<br />

N<br />

- e -<br />

.<br />

...<br />

+ e -<br />

+ e -<br />

+<br />

N CH3 +<br />

N CH3 N<br />

CH 3<br />

CH 3<br />

1,1'-dimethyl-4,4'-dihydrobipyridyl<br />

Fig. 7 – Intermediate resonance structures <strong>of</strong> <strong>paraquat</strong> and full reduction.


__________________________________________________Part I - General Introduction<br />

3. TOXICOKINETICS OF PARAQUAT<br />

The toxicokinetics <strong>of</strong> PQ has been studied in a variety <strong>of</strong> animal species,<br />

especially dogs, rats, and rabbits (Murray and Gibson, 1972; Hawksworth et al., 1981;<br />

Yonemitsu, 1986). The dog seems to be <strong>the</strong> most similar model <strong>of</strong> PQ to human<br />

toxicokinetics (Hawksworth et al., 1981).<br />

3.1 Absorption<br />

Nearly all PQ poisonings result from ingestion. PQ is known to be very rapidly<br />

absorbed, apparently associated with <strong>the</strong> carrier-mediated transport system for choline<br />

on <strong>the</strong> brush-border membrane, thought this absorption from <strong>the</strong> gastrointestinal tract<br />

(GIT) is low (Nagao et al., 1993a). Absorption occurs primarily in <strong>the</strong> small intestine<br />

(poorly from <strong>the</strong> stomach) and is estimated to be 1-5% in humans over a 1–6 hours<br />

period (Baselt and Cravey, 1989; Houze et al., 1990; Houze et al., 1995). Any recent<br />

food ingestion may decrease <strong>the</strong> amount <strong>of</strong> systemic absorption (Meredith and Vale,<br />

1987; Bismuth et al., 1988). Although <strong>the</strong> plasma peak time (Tmax) is not known with<br />

certainty in humans, PQ may be detected in <strong>the</strong> urine as early as 1 hour after ingestion,<br />

and according to data published by Proudfoot et al. (Proudfoot et al., 1979; Proudfoot,<br />

1995) and Smith (Smith, 1988b), peak concentrations (Cmax) in humans are attained<br />

within 4-and possibly within 2-hours after <strong>into</strong>xication. Smith (Smith et al., 1974a)<br />

reported that after oral administration <strong>of</strong> PQ to rats, plasma concentrations remained<br />

relatively constant for 30 hours. During this period <strong>of</strong> time, concentrations in <strong>the</strong> lung<br />

rose progressively to several times <strong>the</strong> plasma concentration. If during <strong>the</strong> first 30 hours,<br />

plasma PQ concentrations were severely reduced by decreasing absorption <strong>of</strong> <strong>the</strong><br />

herbicide from GIT or increasing its elimination by extracorporeal techniques from <strong>the</strong><br />

plasma, lethal concentrations woudn’t reach <strong>the</strong> lungs (Smith et al., 1974a). These<br />

authors also concluded that not only <strong>the</strong> Cmax is responsible for determining <strong>the</strong> lung<br />

levels but also <strong>the</strong> maintenance <strong>of</strong> plasma levels from which <strong>the</strong> lung can take large<br />

amounts <strong>of</strong> PQ. The maintenance <strong>of</strong> such plasma concentrations in <strong>the</strong> rat has been<br />

shown to be <strong>the</strong> result <strong>of</strong> continued PQ absorption from <strong>the</strong> GIT over <strong>the</strong> first 30 hours<br />

after oral administration. Absorption <strong>of</strong> PQ from <strong>the</strong> GIT <strong>into</strong> <strong>the</strong> human bloodstream is<br />

17


Part I - General Introduction __________________________________________________<br />

quite different from that seen in rats; concentrations declined rapidly over <strong>the</strong> first 15<br />

hours after Tmax to much lower levels than those described in rats due to tissue<br />

distribution, and more slowly <strong>the</strong>reafter (Smith, 1987). Thus, in humans, if adsorbents<br />

are to be effective in preventing PQ from entering <strong>the</strong> blood and consequently <strong>the</strong> lung,<br />

<strong>the</strong>y must be administered within a few hours, or accordingly to Bismuth et al.<br />

(Bismuth et al., 1988), within <strong>the</strong> first few minutes after ingestion.<br />

18<br />

Daniel and Gage (Daniel and Gage, 1966) studied <strong>the</strong> absorption <strong>of</strong> [ 14 C]-PQ<br />

following oral and subcutaneous (s.c.) single-dose administration to rats. About 76-90%<br />

<strong>of</strong> <strong>the</strong> oral doses were found in <strong>the</strong> faeces, and 11-20% in <strong>the</strong> urine; most <strong>of</strong> <strong>the</strong> s.c.<br />

dose (73-88%) was found in <strong>the</strong> urine and only 2-14.2% in <strong>the</strong> faeces. These values bear<br />

no relation to <strong>the</strong>ir respective LD50 values (Conning et al., 1969). These studies<br />

evidenced that PQ was poorly absorbed from <strong>the</strong> gut. Rats, guinea-pigs, and monkeys<br />

orally administered LD50 doses <strong>of</strong> [ 14 C]-PQ had low peak plasma concentrations (2.1-<br />

4.8 mg/L) (Murray and Gibson, 1972). Extensive caustic injury to <strong>the</strong> GIT may increase<br />

<strong>the</strong> amount absorbed. The highest concentration was found 1 to 6 hours after an oral<br />

dose depending upon <strong>the</strong> species used (Murray and Gibson, 1972).<br />

Although almost all fatal exposures have resulted from <strong>the</strong> ingestion <strong>of</strong> PQ, a few<br />

case reports have involved ra<strong>the</strong>r extensive skin contamination (Samman and Johnston,<br />

1969; Hearn and Keir, 1971; Vale et al., 1987; Smith, 1988a; H<strong>of</strong>fer and Taitelman,<br />

1989). PQ absorption through animal and human skin was studied in vitro (Walker et<br />

al., 1983). Human skin was shown to be impermeable to PQ, having a very low<br />

permeability coefficient <strong>of</strong> 0.73. Fur<strong>the</strong>rmore, human skin was found to be at least 40<br />

times less permeable than <strong>the</strong> animal skins tested (including rat, rabbit, and guinea-pig)<br />

(Walker et al., 1983). A study <strong>of</strong> <strong>the</strong> percutaneous absorption <strong>of</strong> PQ was undertaken in<br />

six human volunteers by Wester et al. (Wester et al., 1984). It was observed that only<br />

minute quantities <strong>of</strong> PQ were absorbed through intact human skin over 24 hours and<br />

that <strong>the</strong>re was little difference among skin tested at different body sites in its ability to<br />

absorb PQ.<br />

Fatal cases <strong>of</strong> s.c., intravenous (i.v.), intramuscular or intraperitoneal (i.p) injection<br />

<strong>of</strong> PQ were also reported, <strong>the</strong> doses being considerably lower than <strong>the</strong> lethal dose by<br />

ingestion route (Almog and Tal, 1967; Vale et al., 1987; Hsu et al., 2003).<br />

Ocular exposure may cause local caustic injury with ulceration and scarring likely<br />

resulting in a delayed slough <strong>of</strong> corneal epi<strong>the</strong>lium 12–24 hours after exposure, but not<br />

resulting in systemic toxicity (Cant and Lewis, 1968; McKeag et al., 2002).


__________________________________________________Part I - General Introduction<br />

Inhalation <strong>of</strong> PQ used in an agricultural/occupational setting does not allow<br />

sufficient absorption to cause systemic disease, because <strong>of</strong> droplet size (greater than 5<br />

µm) that prevents deep lung exposure and absorption, low product vapour pressure, and<br />

low application concentration (Howard, 1983; Chester and Ward, 1984).<br />

Notwithstanding no fatal cases have been reported from inhalation <strong>of</strong> PQ vapor or<br />

aerosols, toxicity has occurred from this route <strong>of</strong> exposure, since inhalation <strong>of</strong> PQ<br />

droplets may produce nasal and tracheobronchial irritation. An interesting episode in <strong>the</strong><br />

history <strong>of</strong> <strong>the</strong> war against illicit marijuana use, in which large quantities <strong>of</strong> PQ were<br />

sprayed over culture fields in <strong>the</strong> early 1970s is described. By <strong>the</strong>n, PQ was <strong>the</strong><br />

herbicide <strong>of</strong> choice during aerial spraying <strong>of</strong> marijuana by <strong>the</strong> U.S.A. and Mexican<br />

governments. However, after spraying, growers simply harvested <strong>the</strong> crops before <strong>the</strong><br />

plants were exposed to enough sunlight to damage <strong>the</strong> plants, resulting in an apparently<br />

healthy harvest though contaminated with PQ. Concerns regarding <strong>the</strong> smoking <strong>of</strong> PQ-<br />

sprayed marijuana in <strong>the</strong> early 1970s has proved unfounded, because PQ is destroyed by<br />

pyrolysis <strong>into</strong> a relatively nontoxic compound (4,4’-bipyridyl) during <strong>the</strong> smoking<br />

process (Groce and Kimbrough, 1982; Landrigan et al., 1983).<br />

A fatal case <strong>of</strong> PQ absorbed per vagina <strong>of</strong> a 28-year-old woman (who inserted a<br />

tampon inadvertently soaked in PQ) as a consequence <strong>of</strong> respiratory, renal and hepatic<br />

dysfunction was reported (Ong and Glew, 1989).<br />

3.2 Distribution<br />

Despite numerous studies, <strong>the</strong> distribution <strong>of</strong> PQ through <strong>the</strong> different tissues is<br />

still unclear. Dey et al. (Dey et al., 1990) studied <strong>the</strong> toxicokinetics <strong>of</strong> [ 14 C]-PQ in rats<br />

exposed to a single s.c. injection. PQ was rapidly absorbed with a Tmax <strong>of</strong> 20 min. The<br />

toxicokinetic was best characterized by a two-compartment open model, <strong>the</strong> mean t1/2<br />

being approximately 40 hours. Peak concentrations in <strong>the</strong> kidney and lung tissues were<br />

at around 40 min. Never<strong>the</strong>less, <strong>the</strong> majority <strong>of</strong> <strong>the</strong> authors agree that <strong>the</strong> kinetics <strong>of</strong> PQ<br />

in <strong>the</strong> plasma is better described by a three-compartment open model. Murray and<br />

Gibson described a triexponential disappearance <strong>of</strong> [ 14 C]-PQ from <strong>the</strong> plasma after oral<br />

administration (126 mg/Kg) in rats, guinea pigs and monkeys (Murray and Gibson,<br />

1972). The toxicokinetics <strong>of</strong> PQ appears to be similar in human and dog (Hawksworth<br />

et al., 1981; Vandenbogaerde et al., 1984). Hawksworth et al. (Hawksworth et al.,<br />

19


Part I - General Introduction __________________________________________________<br />

1981) described a plasma-concentration-time curve with a triexponential decline, in<br />

dogs, suggesting a three-compartment model:<br />

-Blood is assumed to be <strong>the</strong> central compartment. The concentrations found in<br />

plasma and erythrocytes are approximately <strong>the</strong> same at least in <strong>the</strong> rat (Sharp et al.,<br />

1972);<br />

-The shallow compartment is though to be composed <strong>of</strong> highly perfused tissues<br />

such as <strong>the</strong> kidney, liver, heart, etc. Rapid exchanges occur between this compartment<br />

and blood. The anatomy and physiology <strong>of</strong> <strong>the</strong> lung (a highly vascularized tissue)<br />

suggests <strong>the</strong>refore early exposure to any PQ circulating in <strong>the</strong> blood (Bismuth et al.,<br />

1987);<br />

-The third compartment lies within <strong>the</strong> lungs, especially <strong>the</strong> pneumocytes type I<br />

and II and Clara cells where exchanges with <strong>the</strong> central compartment are slow (Bismuth<br />

et al., 1987). The initial t1/2 <strong>of</strong> PQ in <strong>the</strong> lung was much greater than <strong>the</strong> t1/2 in o<strong>the</strong>r<br />

vital <strong>organ</strong>s (e.g. kidney, liver, muscle, adrenal, spleen, heart, testis), explaining <strong>the</strong><br />

highest PQ lung accumulation (Sharp et al., 1972). The kinetics <strong>of</strong> PQ in <strong>the</strong> rat lungs<br />

shows a rapid decline with an elimination t1/2 <strong>of</strong> 20 min, followed a slower decline with<br />

a t1/2 <strong>of</strong> about 50 hours (Sharp et al., 1972). Peak concentration in lungs is reached 4-5<br />

hours after i.v. administration, and 5-7 hours after ingestion, provided that renal<br />

function is normal (Sharp et al., 1972). Lethal concentrations may be achieved in <strong>the</strong><br />

lung within 6 hours <strong>of</strong> ingestion <strong>of</strong> 35 mg/Kg (Houze et al., 1995). Patients are only<br />

rarely admitted to an experienced hospital in <strong>the</strong> treatment <strong>of</strong> PQ poisonings before <strong>the</strong><br />

pulmonary peak. In <strong>the</strong> presence <strong>of</strong> renal failure (which normally occurs when more<br />

than 20 mg/Kg PQ are ingested), peak pulmonary concentration is not achieved for 15-<br />

20 hours, and may reach very high values (120 hours or longer) (Hawksworth et al.,<br />

1981; Bismuth et al., 1987). Renal failure precludes elimination <strong>of</strong> PQ by its normal<br />

route. Fur<strong>the</strong>rmore, Hawksworth et al. (Hawksworth et al., 1981) suggested that an<br />

impairment <strong>of</strong> renal function by as little as 5% produces a five-fold higher concentration<br />

<strong>of</strong> <strong>the</strong> herbicide in <strong>the</strong> plasma. A critical plasma threshold is needed for active<br />

pulmonary uptake to occur (Manabe and Ogata, 1987). With time, however, it was<br />

shown that <strong>the</strong> concentration in <strong>the</strong> lung did fall to below that in muscle (due to <strong>the</strong><br />

secondary t1/2 in <strong>the</strong> muscle). Considering that muscle represents a large percentage <strong>of</strong><br />

<strong>the</strong> body mass, it may be considered an important reservoir <strong>of</strong> PQ (Murray and Gibson,<br />

1972; Sharp et al., 1972).<br />

20


__________________________________________________Part I - General Introduction<br />

Houze et al. (Houze et al., 1990) studied <strong>the</strong> toxicokinetics <strong>of</strong> PQ in 18 cases <strong>of</strong><br />

acute human poisoning. The concentration-time course was described by using a<br />

biexponential curve suggesting a two-compartment model with absorption, distribution<br />

and elimination phases. Plasma PQ concentration exhibited a mean distribution half-life<br />

(t1/2α) <strong>of</strong> 5 hours and a mean elimination half-life (t1/2β) <strong>of</strong> 84 hours. Tissue PQ<br />

distribution was ubiquitous with an apparent volume <strong>of</strong> distribution ranging from 1.2 to<br />

1.6 L/Kg. Muscle represented an important reservoir explaining <strong>the</strong> long persistence <strong>of</strong><br />

PQ in plasma and urine for several weeks or months after poisoning (Smith, 1988b).<br />

The volume <strong>of</strong> distribution <strong>of</strong> PQ estimated from a kinetic study in one patient was 2.75<br />

L/Kg (Davies, 1987). Recently, immunohistochemical studies were used to demonstrate<br />

<strong>the</strong> distribution and localization <strong>of</strong> PQ in several <strong>organ</strong>s. In <strong>the</strong> skin, PQ was localized<br />

in <strong>the</strong> ducts <strong>of</strong> sweat glands and sebaceous glands between 3 and 10 days after PQ i.v.<br />

injection (Nagao et al., 1993c). In <strong>the</strong> eyes, weak positive findings were observed in<br />

nerve fibers <strong>of</strong> retina between 3 and 10 days after <strong>the</strong> injection. In <strong>the</strong> cornea, PQ was<br />

localized in epi<strong>the</strong>lial cells at <strong>the</strong> first 3 hours and between 3 and 10 days after PQ<br />

administration. Since skin occupies a vast area <strong>of</strong> <strong>the</strong> body in animals, as an <strong>organ</strong>, it<br />

seems to be an important storage pool for <strong>the</strong> redistribution <strong>of</strong> PQ (Nagao et al., 1993c).<br />

PQ was also found in immune and haematopoietic systems (Nagao et al., 1994). In <strong>the</strong><br />

bone marrow, PQ was localized in several types <strong>of</strong> blood cells (granulocyte, erythrocyte<br />

and megakaryocyte) and <strong>the</strong>ir precursors between 24 hours and 7 days after <strong>the</strong> i.v.<br />

administration. In <strong>the</strong> thymus, PQ was mainly localized in <strong>the</strong> medulla between 12<br />

hours and 7 days after administration, whereas in <strong>the</strong> spleen, it was mainly localized in<br />

<strong>the</strong> red pulp between 12 hours and 10 days after administration <strong>of</strong> PQ (Nagao et al.,<br />

1994). In <strong>the</strong> stomach, PQ was localized in <strong>the</strong> epi<strong>the</strong>lial cells between 24 hours and 10<br />

days after administration, whereas in <strong>the</strong> esophagus, PQ was localized in epi<strong>the</strong>lial cells<br />

and <strong>the</strong> lamina propria mucosa between 12 hours and 10 days after administration.<br />

Although <strong>the</strong>se findings were similar to those observed in <strong>the</strong> intestine <strong>of</strong> rats, no clear<br />

changes in <strong>the</strong> distribution <strong>of</strong> PQ with time were observed, suggesting that <strong>the</strong> stomach<br />

and esophagus are important reservoirs for <strong>the</strong> redistribution <strong>of</strong> PQ (Nagao et al.,<br />

1993b).<br />

Concerning <strong>the</strong> PQ binding to plasma proteins, controversial data exist. For many<br />

years PQ was thought not to bind to plasma proteins (Lock and Ishmael, 1979).<br />

Recently Jaiswal et al. (Jaiswal et al., 2002) showed <strong>the</strong> binding <strong>of</strong> PQ to plasma<br />

albumin by using a fluorescence technique.<br />

21


Part I - General Introduction __________________________________________________<br />

3.2.1 Preferential accumulation in <strong>the</strong> lung<br />

22<br />

Irrespective <strong>of</strong> <strong>the</strong> route <strong>of</strong> administration, <strong>the</strong> lung and <strong>the</strong> kidney are <strong>the</strong> <strong>organ</strong>s<br />

showing <strong>the</strong> highest concentrations <strong>of</strong> PQ (Murray and Gibson, 1972; Sharp et al.,<br />

1972; Ilett et al., 1974). The distribution <strong>of</strong> [ 14 C]PQ <strong>into</strong> various tissues after oral<br />

administration <strong>of</strong> 680 mol/Kg to rats was followed as a function <strong>of</strong> time by Rose et al.<br />

(Rose et al., 1976a). They showed that although <strong>the</strong> plasma concentration <strong>of</strong> PQ<br />

remained constant between 2 and 30 hours after administration, PQ concentrations in<br />

<strong>the</strong> lung exhibited a time-dependent increase over <strong>the</strong> same period. None <strong>of</strong> <strong>the</strong> o<strong>the</strong>r<br />

studied <strong>organ</strong>s showed this time-dependent accumulation. Rose et al. (Rose et al., 1974)<br />

also demonstrated that slices <strong>of</strong> lung incubated with [ 14 C]PQ exhibited a time-dependent<br />

accumulation <strong>of</strong> radioactivity. In addition, lung slices were <strong>the</strong> only tissue slices in<br />

which PQ accumulated at a concentration significantly higher than that in <strong>the</strong> medium.<br />

These studies demonstrated that lung, and no o<strong>the</strong>r major tissue, is able to accumulate<br />

PQ against a concentration gradient. After in vivo administration, PQ levels in <strong>the</strong><br />

kidney did not show a time-dependent increase, but were never<strong>the</strong>less higher than those<br />

in <strong>the</strong> lung throughout <strong>the</strong> first 30 hours (Rose et al., 1976a). These high concentrations<br />

<strong>of</strong> PQ in <strong>the</strong> kidney probably result from <strong>the</strong> fact that this <strong>organ</strong> represents <strong>the</strong><br />

predominant route <strong>of</strong> elimination <strong>of</strong> PQ from <strong>the</strong> circulation and are likely to constitute<br />

extracellular ra<strong>the</strong>r than intracellular PQ. They may also underlie <strong>the</strong> observation that<br />

renal failure <strong>of</strong>ten occurs in PQ poisoning, especially during early stages. Taken<br />

toge<strong>the</strong>r, <strong>the</strong>se studies strongly suggest that <strong>the</strong> lungs are a specific target for <strong>the</strong><br />

pathological effects <strong>of</strong> PQ because <strong>of</strong> its selective accumulation by this <strong>organ</strong>. PQ<br />

pulmonary concentrations can be 6 to 10 times higher than those in <strong>the</strong> plasma, and <strong>the</strong><br />

compound is retained in <strong>the</strong> lung even when blood levels are starting to decrease.<br />

3.2.2 Lung accumulation through <strong>the</strong> polyamine uptake system<br />

Early work by Rose et al. (Rose et al., 1974) demonstrated that <strong>the</strong> accumulation<br />

<strong>of</strong> PQ <strong>into</strong> rat lung slices occurred against a concentration gradient and could be<br />

abolished by metabolic inhibitors such as cyanide or rotenone, suggesting that <strong>the</strong><br />

uptake is an adenosine triphosphate (ATP)-driven process. The accumulation also<br />

exhibited saturation kinetics with an apparent Michaelis-Menten constant (Km) <strong>of</strong> 70 µM


__________________________________________________Part I - General Introduction<br />

and a maximal rate (Vmax) <strong>of</strong> 300 nmol PQ × g wet weight 1 × h 1 . These observations,<br />

coupled with findings that PQ is nei<strong>the</strong>r metabolized by <strong>the</strong> lung (Conning et al., 1969;<br />

Ilett et al., 1974) nor becomes covalently bound to any degree (Ilett et al., 1974;<br />

Sullivan and Montgomery, 1983), suggest that its accumulation is mediated through<br />

binding to, and subsequent translocation <strong>into</strong> cells by a carrier system. Active<br />

accumulation <strong>of</strong> PQ via transport systems exhibiting similar kinetic parameters (Table<br />

4) was also demonstrated in lung slices taken from o<strong>the</strong>r species [beagle dogs, New<br />

Zealand white rabbits, and cynomolgus monkeys (Macaca fascicularis)], including<br />

humans (Rose et al., 1976a). The kinetic constants for human and rat lung are<br />

statistically similar, suggesting that <strong>the</strong> rat may be a good experimental model for <strong>the</strong><br />

study <strong>of</strong> PQ accumulation in <strong>the</strong> human lung. Thus, it seems likely that <strong>the</strong> human lung<br />

does possess a similar transport to that characterized in <strong>the</strong> rat lung and this process<br />

accounts for <strong>the</strong> selective accumulation and hence <strong>the</strong> selective toxicity <strong>of</strong> PQ to <strong>the</strong><br />

human lung.<br />

Table 4 - Kinetic constants for <strong>the</strong> accumulation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong> lung tissue slices from<br />

various species Adapted from Rose et al. (Rose et al., 1974).<br />

Species Km (µM)<br />

Vmax<br />

(nmol <strong>of</strong> PQ/g<br />

tissue/hour)<br />

Rat 70 300<br />

Mouse 68 556<br />

Syrian hamster 77 452<br />

Guinea-pig 96 49<br />

Rabbit 0.05 20<br />

Man 40 300<br />

Monkey 70 50<br />

Dog 60 10<br />

Carrier-mediate PQ uptake also occurs in <strong>the</strong> isolated perfused lung, <strong>into</strong> which<br />

accumulation <strong>of</strong> PQ to a concentration in excess <strong>of</strong> that in <strong>the</strong> perfusate has been<br />

observed (Rannels et al., 1985). However, <strong>the</strong> kinetics <strong>of</strong> this process appears somewhat<br />

different in comparison to <strong>the</strong> lung slices. The onset <strong>of</strong> active transport is preceded by<br />

an initial lag phase during which <strong>the</strong> intracellular PQ concentration approaches that in<br />

<strong>the</strong> perfusate, possibly because <strong>the</strong> endo<strong>the</strong>lium functions as a barrier between <strong>the</strong> intravascular<br />

and <strong>the</strong> interstitial compartments. Only when concentrations proximal to <strong>the</strong><br />

23


Part I - General Introduction __________________________________________________<br />

epi<strong>the</strong>lium have risen sufficiently (relative to <strong>the</strong> Km value for its uptake) would active<br />

accumulation occur at a significant rate. Since in <strong>the</strong> isolated perfused lung, <strong>the</strong> delivery<br />

<strong>of</strong> PQ occurs through <strong>the</strong> vasculature, <strong>the</strong> sequence or pattern <strong>of</strong> exposure <strong>of</strong> lung cells<br />

to PQ in this system may more closely resemble that occurring in vivo comparatively to<br />

<strong>the</strong> lung slice model, in which <strong>the</strong> epi<strong>the</strong>lium becomes directly exposed. The<br />

observation that in vivo (Smith and Heath, 1974) <strong>the</strong> rate <strong>of</strong> accumulation (Vmax) <strong>of</strong> PQ<br />

in <strong>the</strong> lung was only one-seventh <strong>of</strong> that found in vitro in lung slices led to a search for<br />

compounds present in plasma and capable <strong>of</strong> blocking <strong>the</strong> uptake <strong>of</strong> PQ in <strong>the</strong> lung<br />

(Lock et al., 1976). Subsequent to <strong>the</strong> identification <strong>of</strong> this transport system, a number<br />

<strong>of</strong> naturally occurring amines have been identified, which competitively inhibit <strong>the</strong><br />

uptake <strong>of</strong> PQ <strong>into</strong> lung tissue and which <strong>the</strong>mselves act as substrates and accumulated<br />

in rat lung slices in a saturable manner, obeying Michaelis-Menten kinetics. These<br />

amines include <strong>the</strong> diamines putrescine and cadaverine, <strong>the</strong> oligoamines spermidine and<br />

spermine (Smith, 1982; Wyatt et al., 1988), and <strong>the</strong> disulfide cystamine (Lewis et al.,<br />

1989) (Fig. 8). An important property <strong>of</strong> <strong>the</strong>se specific polyamines is that <strong>the</strong>y are<br />

positively charged at a physiological pH, and, consequently, <strong>the</strong>y have a high affinity<br />

toward negatively charged cellular molecules. Thus polyamines are very soluble in<br />

water, and <strong>the</strong>y exert strong cation-anion interactions with macromolecules, mainly with<br />

DNA and RNA (Marczynski, 1985), a feature that represents <strong>the</strong>ir best-known direct<br />

physiological role in cellular functions such as cell growth, division, and differentiation<br />

(Janne et al., 1978; Heby, 1981).<br />

24


__________________________________________________Part I - General Introduction<br />

N<br />

H 2<br />

C<br />

H 3<br />

N<br />

H 2<br />

N<br />

H<br />

+<br />

N<br />

N<br />

H 2<br />

N<br />

H 2<br />

0.702 nm<br />

0.622 nm<br />

N<br />

H<br />

H<br />

N<br />

+<br />

N CH3 Fig. 8 - Chemical structure <strong>of</strong> <strong>paraquat</strong> (A) and putrescine (B), showing geometric<br />

standards <strong>of</strong> <strong>the</strong> distance between N atoms (optimal distance to fit polyamine uptake<br />

system is unknown). The chemical structure <strong>of</strong> cadaverin (C), spermidine (D) and<br />

spermine (E) is also presented<br />

A possible gene coding for a polyamine transporter (TPO1) was isolated from<br />

eukaryotic cells and introduced <strong>into</strong> yeast cells (McNemar et al., 2001). Yeast cells<br />

heterologously expressing TPO1, become sensitive to polyamines. For mammals,<br />

though, it is not known whe<strong>the</strong>r <strong>the</strong> transporter(s) is (are) located at <strong>the</strong> apical or basal<br />

side <strong>of</strong> <strong>the</strong> cells, and how <strong>the</strong> expression <strong>of</strong> <strong>the</strong> gene is regulated. The suggestion has<br />

been put forward that <strong>the</strong> polyamines, which, as above mentioned, are known to<br />

regulate cell growth, may play a role in <strong>the</strong> differentiation <strong>of</strong> alveolar epi<strong>the</strong>lial type II<br />

cells to type I cells (Smith, 1982). It has also been proposed that cystamine represents a<br />

source <strong>of</strong> taurine, which may have an antioxidant role in <strong>the</strong> lung (Lewis et al., 1989).<br />

Subjects in whom acute fulminant poisoning occurs generally ingested more than<br />

40 mg <strong>of</strong> PQ ion/Kg <strong>of</strong> body weight (BW) (Vale et al., 1987). In <strong>the</strong>se cases, <strong>the</strong> role <strong>of</strong><br />

<strong>the</strong> pulmonary transport system clearly has a negligible role on <strong>the</strong> evolution <strong>of</strong> <strong>the</strong><br />

NH 2<br />

NH 2<br />

NH 2<br />

NH 2<br />

A<br />

B<br />

C<br />

D<br />

E<br />

25


Part I - General Introduction __________________________________________________<br />

<strong>into</strong>xication, which progresses to multi<strong>organ</strong> system failure. However, in cases <strong>of</strong><br />

moderate poisoning, where PQ plasma concentrations appear to be at <strong>the</strong> order <strong>of</strong> 10 to<br />

20 μM, kinetic considerations suggest that only cells actively accumulating PQ would<br />

achieve <strong>the</strong> intracellular concentrations necessary to cause significant cellular damage.<br />

3.2.3 Structural requirements for <strong>the</strong> pulmonary polyamine uptake system<br />

An important aim <strong>of</strong> earlier studies concerning pulmonary polyamine uptake<br />

system (PUS) was to discover <strong>the</strong> structural requirements for substrates <strong>of</strong> <strong>the</strong> transport<br />

system in order to find possible antagonists capable <strong>of</strong> preventing PQ from entering its<br />

target cells. Ross and Krieger (Ross and Krieger, 1981) established that to act as a<br />

substrate for <strong>the</strong> pulmonary PUS, a molecule must possess <strong>the</strong> following characteristics:<br />

i) two or more positively charged nitrogen atoms, ii) maximum positivity <strong>of</strong> charge<br />

surrounding <strong>the</strong>se nitrogens, iii) a nonpolar group between <strong>the</strong>se charges, and iv) a<br />

minimum <strong>of</strong> steric hindrance. Gordonsmith et al. (Gordonsmith et al., 1983) have<br />

demonstrated that <strong>the</strong> optimum distance (essential for binding and consequently, for<br />

transport) between <strong>the</strong> nitrogen centers is four methylene groups (about 0.622 nm as it<br />

occurs in putrescine), although a spacing between four and seven methylene groups is<br />

tolerated. These assumptions explain how polyamines and PQ (with ≈ 0.702 between<br />

two positively charged nitrogens) can share a common uptake system, but also why PQ<br />

(with its steric hindrance <strong>of</strong> <strong>the</strong> nitrogens by <strong>the</strong> pyridine rings) is a less successful<br />

substrate (Smith, 1987). The affinity <strong>of</strong> <strong>the</strong> uptake system for <strong>the</strong> polyamines appeared<br />

to be sevenfold higher (i.e., exhibiting a lower apparent Km) than that <strong>of</strong> PQ (Smith,<br />

1982). Although PQ proved to be a ra<strong>the</strong>r "poor" substrate (higher Km than polyamines)<br />

for <strong>the</strong> PUS, it is undoubtedly "recognized" as a substrate, probably as a consequence <strong>of</strong><br />

its structural similarity to <strong>the</strong>se endogenous substrates (Fig. 8), and is <strong>the</strong>refore<br />

mistakenly accumulated <strong>into</strong> <strong>the</strong> lung, especially in <strong>the</strong> alveolar type I and II cells and in<br />

<strong>the</strong> Clara cells, through this transport pathway (Smith, 1982). Later, O'Sullivan et al.<br />

(O'Sullivan et al., 1991) showed that many putrescine analogues competitively inhibit<br />

putrescine and PQ uptake. The authors established that <strong>the</strong> inhibition <strong>of</strong> putrescine<br />

uptake by analogs decreases with increasing N-alkylation and those analogues with a<br />

bulky substituent <strong>of</strong> <strong>the</strong> butyl chain do not inhibit <strong>the</strong> uptake at all. The strongest<br />

inhibition was found with N-(4-aminobutyl)aziridine; this cytotoxic compound does not<br />

26


__________________________________________________Part I - General Introduction<br />

seem to alter <strong>the</strong> polyamine Vmax but might fit <strong>into</strong> <strong>the</strong> substrate binding site <strong>of</strong> <strong>the</strong><br />

receptor. The selective accumulation/retention <strong>of</strong> PQ in lung tissue provides a plausible<br />

explanation for this <strong>organ</strong> selectivity to damage in comparison with o<strong>the</strong>r tissues.<br />

Although <strong>the</strong> disposition <strong>of</strong> PQ in human tissues has not been as extensively studied as<br />

in experimental animals, <strong>the</strong> major <strong>organ</strong>s affected in man are also <strong>the</strong> lung and kidney.<br />

Therefore, it seems likely that PQ is selectively accumulated in <strong>the</strong> human lung and<br />

excreted by <strong>the</strong> kidney.<br />

DQ exposure produce signs and symptoms similar to those <strong>of</strong> PQ except for one<br />

important system - <strong>the</strong> pulmonary system (Jones and Vale, 2000). In contrast to PQ, DQ<br />

is not a substrate for <strong>the</strong> pulmonary PUS and <strong>the</strong>refore is not selectively pneumotoxic.<br />

In fact DQ exhibits a much smaller intramolecular distance between <strong>the</strong> two charged<br />

nitrogen atoms, explaining its much greater safety margin (Rose and Smith, 1977).<br />

3.2.4 Characterization <strong>of</strong> <strong>the</strong> pulmonary polyamine uptake system<br />

It is clear that <strong>the</strong>re will be a range <strong>of</strong> endogenous and exogenous compounds that<br />

are capable <strong>of</strong> using this uptake system. Smith and Wyatt (Smith and Wyatt, 1981) and<br />

Lewis et al. (Lewis et al., 1989) showed that <strong>the</strong> uptake <strong>of</strong> putrescine and cystamine in<br />

rat lung slices was not dependent on <strong>the</strong> sodium (Na + ) concentration in <strong>the</strong> medium. In<br />

contrast to <strong>the</strong>se observations, Rannels et al. (Rannels et al., 1989) found that, in type II<br />

pneumocytes <strong>the</strong> uptake <strong>of</strong> putrescine and spermidine was dependent on Na + , whereas<br />

spermine uptake was not, indicating that polyamine uptake may take place via different<br />

uptake systems. However, in <strong>the</strong>se experiments, <strong>the</strong> nature and concentration <strong>of</strong> <strong>the</strong> ions<br />

used to replace Na + were probably critical factors, because it has been shown that a<br />

supplement <strong>of</strong> NaCl, LiCl, or choline significantly reduced <strong>the</strong> uptake <strong>of</strong> polyamines<br />

due to <strong>the</strong> increase <strong>of</strong> osmotic pressure (Rannels et al., 1989). On <strong>the</strong> o<strong>the</strong>r hand,<br />

Kumagai and Johnson (Kumagai and Johnson, 1988) showed that replacement <strong>of</strong> Na + by<br />

mannitol or sucrose did not modulate putrescine uptake in rat enterocytes, whereas<br />

replacement by choline, lithium (Li + ), N-methyl-D-glucamine, or tetramethylammonium<br />

did. It was hypo<strong>the</strong>sized that cations can interact with <strong>the</strong> carrier but that no co-transport<br />

<strong>of</strong> Na + is involved in putrescine uptake. Ano<strong>the</strong>r issue is whe<strong>the</strong>r <strong>the</strong>re is one or more<br />

pulmonary PUSs. In bovine arterial smooth muscle cells, Aziz et al. (Aziz et al., 1994)<br />

and Jänne et al. (Janne et al., 1978) found that putrescine was accumulated through an<br />

27


Part I - General Introduction __________________________________________________<br />

uptake system that is also used by spermidine, spermine, PQ, and methylglyoxal bis-<br />

(guanylhydrazone) (MGBG), but spermidine and spermine were also accumulated<br />

through a different uptake system insensitive to putrescine and PQ, and only partially<br />

sensitive to <strong>the</strong> presence <strong>of</strong> MGBG. Similarly, one study using suspensions <strong>of</strong> freshly<br />

isolated type II pneumocytes (Chen et al., 1992) showed that putrescine uptake was<br />

inhibited by PQ (and vice versa) in a partially competitive manner. These authors<br />

postulated that <strong>the</strong> PUS in type II cells for PQ and putrescine possessed two separate<br />

sites, one for each substrate, and that binding at one site leads to a conformational<br />

change in <strong>the</strong> o<strong>the</strong>r. However, such partially competitive inhibition was not found in<br />

o<strong>the</strong>r studies using hamster (Hoet et al., 1995) or human (Hoet et al., 1994) type II<br />

pneumocytes. In ano<strong>the</strong>r study performed in rat lung slices <strong>the</strong> inhibition <strong>of</strong> PQ<br />

accumulation in presence <strong>of</strong> putrescine resulted from a process that appears to be<br />

competitive (Karl and Friedman, 1983).<br />

3.2.5 Cellular localization <strong>of</strong> <strong>the</strong> polyamine uptake system in <strong>the</strong> lung<br />

The problem <strong>of</strong> <strong>the</strong> localization <strong>of</strong> <strong>the</strong> PUS in <strong>the</strong> lung was addressed first by<br />

identifying <strong>the</strong> cellular targets for <strong>the</strong> toxicity <strong>of</strong> PQ and later by identifying <strong>the</strong> site <strong>of</strong><br />

accumulation <strong>of</strong> radiolabeled PQ and/or polyamines. Smith and Wyatt (Smith and<br />

Wyatt, 1981) performed morphological and functional studies to localize <strong>the</strong> site <strong>of</strong><br />

cytotoxicity <strong>of</strong> PQ in lung slices taken from PQ (20 mg/Kg)-exposed rats. Lung slices<br />

taken from rats 24 hours after treatment evidenced morphological damage to type I and<br />

type II cells and <strong>the</strong>ir ability to take up putrescine (10 µM) or PQ (10 µM) was<br />

impaired, thus suggesting that type I or type II pneumocytes are <strong>the</strong> site <strong>of</strong> uptake <strong>of</strong><br />

putrescine and PQ. Ano<strong>the</strong>r experimental approach to determine <strong>the</strong> site <strong>of</strong> polyamine<br />

uptake resulted from autoradiography. Waddell and Marlowe (Waddell and Marlowe,<br />

1980) showed that after <strong>the</strong> i.v. administration <strong>of</strong> [ 14 C]PQ (10 µM) to mice, distribution<br />

<strong>of</strong> <strong>the</strong> label corresponded to that in alveolar type II cells. Studies with rat lung slices by<br />

Nemery et al. (Nemery et al., 1987) clearly demonstrated <strong>the</strong> presence <strong>of</strong> [ 3 H]putrescine<br />

in alveolar type II cells and also in bronchiolar Clara cells (Fig. 9).<br />

28


__________________________________________________Part I - General Introduction<br />

Fig. 9 - Autoradiographs <strong>of</strong> rat lung tissue incubated with [ 3 H]putrescine. Resin sections<br />

1 µm thick were stained with toluidine blue and examined by light microscopy.<br />

Labeling occurs in alveolar walls and in alveolar type II pneumocytes (A and B, arrows).<br />

There is no labeling in macrophages (B) or in walls <strong>of</strong> vessels, but Clara cells<br />

(arrowheads) in bronchiolar epi<strong>the</strong>lium (C) show intense labeling. Original<br />

magnifications: ×600 in A; ×1,500 in B and C. Adapted from Nemery et al. (Nemery et<br />

al., 1987).<br />

Wyatt et al. (Wyatt et al., 1988), who carried out both in vivo and in vitro studies,<br />

also showed uptake <strong>of</strong> [ 3 H]PQ, [ 3 H]putrescine, [ 3 H]spermidine, and [ 3 H]spermine by<br />

alveolar type II cells and, at least in vitro, also by Clara cells. Hoet and co-workers also<br />

visualized, by ultrastructural autoradiography, [ 14 C]putrescine in both type I and type II<br />

cells <strong>of</strong> <strong>the</strong> alveolar epi<strong>the</strong>lium, but not over <strong>the</strong> endo<strong>the</strong>lium or any cells <strong>of</strong> <strong>the</strong><br />

29


Part I - General Introduction __________________________________________________<br />

interstitium, in hamster (Hoet et al., 1995) and human (Hoet et al., 1993) lung slices<br />

(Fig. 10).<br />

Fig. 10 - Autoradiographs <strong>of</strong> human lung tissue incubated with 2.5 µM [ 3 H]putrescine.<br />

Unstained resin sections 1 µm thick were examined by electron spectroscopic imaging.<br />

a-d: 4 different alveolar spaces lined with type II and type I pneumocytes. Silver grains<br />

30


__________________________________________________Part I - General Introduction<br />

are evident over type II pneumocytes (long arrows) and lining <strong>of</strong> alveoli (short arrows)<br />

but not over erythrocytes (*) or paranuclear regions <strong>of</strong> endo<strong>the</strong>lium. Cellular and<br />

noncellular components <strong>of</strong> alveolar interstitium were largely devoid <strong>of</strong> silver grains.<br />

Silver grains were uniformly distributed over both nucleus and cytoplasm <strong>of</strong> type II<br />

cells. Bars, 1 µm. Adapted from Hoet et al. (Hoet et al., 1993).<br />

Dinsdale et al. (Dinsdale et al., 1991) also clearly demonstrated labelling in <strong>the</strong><br />

alveolar type I cell in rat by autoradiography at <strong>the</strong> electron-microscopic level. Saunders<br />

et al. (Saunders et al., 1989) suggested that alveolar macrophages were <strong>the</strong> site <strong>of</strong><br />

putrescine and spermidine accumulation in rabbits, a species that shows a different<br />

response to PQ (Smith et al., 1978). Masek and Richards (Masek and Richards, 1990)<br />

demonstrated that <strong>the</strong> toxicity <strong>of</strong> PQ to isolated mouse Clara cells could be decreased by<br />

addition <strong>of</strong> putrescine to <strong>the</strong> incubation medium. However, although this could be due<br />

to <strong>the</strong> inhibition <strong>of</strong> PQ accumulation <strong>into</strong> <strong>the</strong> cells, intracellular PQ levels were not<br />

determined.<br />

The specific distribution <strong>of</strong> <strong>the</strong> PUS in a number <strong>of</strong> individual cell types is <strong>of</strong><br />

considerable importance in attempting to understand <strong>the</strong> mechanism <strong>of</strong> PQ toxicity.<br />

Usually, data describing <strong>the</strong> amount <strong>of</strong> PQ present in <strong>the</strong> lung are expressed on a per<br />

gram wet-weight basis. Since <strong>the</strong>re are more than 40 different cell types in <strong>the</strong> lung<br />

(Sorokin, 1970), each with unique and functional activities, <strong>the</strong> concentration expressed<br />

on this basis will underestimate by perhaps as much as two orders <strong>of</strong> magnitude <strong>the</strong><br />

concentration <strong>of</strong> PQ within specific cell types.<br />

3.3 Metabolism<br />

Only a small fraction <strong>of</strong> orally-administered PQ is metabolized, <strong>the</strong> greater part<br />

being excreted unchanged in <strong>the</strong> urine. Daniel and Cage (Daniel and Gage, 1966)<br />

undertook a study in rats using [ 14 C]-labeled PQ dichloride, and some evidence <strong>of</strong><br />

metabolism by micr<strong>organ</strong>isms in <strong>the</strong> gut, following oral dosing <strong>of</strong> rats, was found. Of<br />

<strong>the</strong> total oral dose <strong>of</strong> PQ, 30% <strong>of</strong> <strong>the</strong> label was present in <strong>the</strong> gut as metabolic products.<br />

Fur<strong>the</strong>rmore, a small amount <strong>of</strong> metabolite was present in <strong>the</strong> urine after oral but not<br />

s.c. administration, suggesting <strong>the</strong> absorption <strong>of</strong> metabolites from <strong>the</strong> gut. Studies in<br />

vitro, using faecal homogenates suggested that microbiological biotransformation was<br />

31


Part I - General Introduction __________________________________________________<br />

responsible for this effect. However, in ano<strong>the</strong>r study, reported by Murray and Gibson<br />

(Murray and Gibson, 1972), gavage administration <strong>of</strong> [ 14 C]-labeled PQ to rats, guinea<br />

pigs, and monkeys, it was not observed any formation <strong>of</strong> metabolites.<br />

3.4 Elimination<br />

According to <strong>the</strong> above comments, PQ is rapidly excreted by <strong>the</strong> kidneys. Daniel<br />

and Cage (Daniel and Gage, 1966) recovered virtually all <strong>of</strong> a PQ oral dose in <strong>the</strong><br />

excreta <strong>of</strong> rats by 2 days. Absorbed PQ is almost completely eliminated unchanged by<br />

<strong>the</strong> renal system (Baselt and Cravey, 1989) and is accomplished by both glomerular<br />

filtration and active tubular secretion. Hawksworth et al. (Hawksworth et al., 1981)<br />

studied <strong>the</strong> elimination <strong>of</strong> PQ in dogs. After an i.v. administration <strong>of</strong> low doses <strong>of</strong><br />

[ 14 C]–labelled PQ (30 to 50 μg/Kg), it was rapidly excreted in <strong>the</strong> urine, with 80-90%<br />

being excreted in <strong>the</strong> first 6 hours and urinary recovery being almost 100% complete by<br />

24 hours. The PQ clearance [CLPQ, (28 mL/min)] was greater than <strong>the</strong> glomerular<br />

filtration rate (GFR), suggesting a process <strong>of</strong> active secretion, which may exceed 200<br />

mL/min when renal function is normal (Bismuth et al., 1987). Tubular secretion was<br />

inhibited by N-methylnicotinamide (NMN) infusion, suggesting that PQ is secreted<br />

through an active transport process with high affinity for alkaline coumpounds<br />

(Hawksworth et al., 1981). After NMN administration CLPQ approximates to creatinine<br />

clearance (CLCr). Following administration <strong>of</strong> large doses <strong>of</strong> PQ (20 mg/Kg), <strong>the</strong> CLPQ<br />

and CLCr decreased due to renal tubular necrosis reducing urinary output and CLPQ by<br />

10 to 20 times after <strong>the</strong> first few hours. Consequently, <strong>the</strong> urinary t1/2 increases<br />

(exceeding 120 hours). Chan et al. (Chan et al., 1997) studied <strong>the</strong> renal clearance <strong>of</strong> PQ<br />

in male Wistar rats using inulin as <strong>the</strong> marker <strong>of</strong> GFR. The obtained results<br />

demonstrated that <strong>the</strong> excretion <strong>of</strong> PQ was greater than <strong>the</strong> GFR, concentration<br />

dependent and saturable, indicating that it was secreted by an active transport system.<br />

The excretion <strong>of</strong> PQ was predominantly dependent on <strong>the</strong> GFR with a small secretory<br />

component (Km = 8.5 ± 3.1 µM, Vmax = 114 ± 19 nmol/Kg per min). The CLPQ was not<br />

inhibited by high doses <strong>of</strong> cimetidine, or p-aminohippurate (PAH). However, quinine<br />

and NMN reduced <strong>the</strong> fractional excretion <strong>of</strong> PQ, suggesting that <strong>the</strong>y share <strong>the</strong> same<br />

cation transport system with PQ. Sharp et al. (Sharp et al., 1972) reported a biphasic<br />

elimination <strong>of</strong> PQ from <strong>the</strong> plasma <strong>of</strong> rats after i.v. administration. The initial rapid<br />

32


__________________________________________________Part I - General Introduction<br />

phase had a 20-30 min t1/2, and <strong>the</strong> slower phase a t1/2 <strong>of</strong> 56 hours. Murray and Gibson<br />

(Murray and Gibson, 1972) also showed prolonged PQ elimination after oral<br />

administration to rats, guinea-pigs, and monkeys. The urinary and faecal routes were<br />

equally important in all species studied. The faecal content was mainly due to<br />

elimination <strong>of</strong> unabsorbed PQ. Prolonged elimination <strong>of</strong> PQ in all tested animals<br />

indicated retention <strong>of</strong> <strong>the</strong> herbicide in <strong>the</strong> body. Despite <strong>the</strong> rapid PQ excretion, <strong>the</strong><br />

kidneys are not very efficient at removing it from blood, since <strong>the</strong>re is considerable<br />

reabsorption <strong>of</strong> PQ through <strong>the</strong> proximal convoluted tubules (Ferguson, 1971). This<br />

reabsorption appears to be a process <strong>of</strong> simple passive diffusion and is <strong>the</strong>refore reduced<br />

by rapid diuresis. This fact has considerable clinical significance. Biliary excretion <strong>of</strong><br />

PQ is small (Daniel and Gage, 1966; Hughes et al., 1973).<br />

Data from <strong>the</strong> limited human studies point to an elimination pattern similar to <strong>the</strong><br />

excretion observed in experimental animals, unchanged PQ elimination being<br />

essentially renal through two pathways: glomerular filtration and tubular secretion<br />

(Bismuth et al., 1988). Tubular reabsorption is minimal (Beebeejaun et al., 1971). With<br />

normal renal function, CLPQ is much greater than CLCr, which enables excretion <strong>of</strong> high<br />

concentrations and large amounts <strong>of</strong> <strong>the</strong> herbicide within <strong>the</strong> first few hours <strong>of</strong><br />

ingestion. Ingestion <strong>of</strong> large doses <strong>of</strong> PQ causes tubular necrosis with a rapid decrease<br />

in <strong>the</strong> GFR and tubular secretion, and <strong>the</strong> consequent increase <strong>of</strong> <strong>the</strong> elimination t1/2<br />

(Bismuth et al., 1987; Bismuth et al., 1988). However, even without renal failure, in<br />

humans, PQ excretion showed to be slower than in animals, since it was detected in <strong>the</strong><br />

urine 7 days after ingestion (Carson, 1972) or as long as 26 days (Beebeejaun et al.,<br />

1971). During this prolonged excretion time <strong>the</strong> concentration <strong>of</strong> PQ in blood was<br />

shown to be below <strong>the</strong> limit <strong>of</strong> detection; tissues act as depots from which PQ is<br />

released at a low rate (Carson, 1972). Never<strong>the</strong>less, in humans, over 90% is excreted<br />

unchanged within 12 to 24 hours <strong>of</strong> ingestion, if renal function remains normal (Houze<br />

et al., 1990). Small amounts <strong>of</strong> PQ have been recovered in <strong>the</strong> bile post-mortem. Thus<br />

enterohepatic recirculation may also exist in humans (Douze et al., 1975).<br />

33


Part I - General Introduction __________________________________________________<br />

34<br />

4. BIOCHEMICAL MECHANISMS OF PARAQUAT TOXICITY<br />

4.1 Mechanism <strong>of</strong> toxicity<br />

A considerable amount <strong>of</strong> work has been done on <strong>the</strong> toxicodynamic <strong>mechanisms</strong><br />

that underlie <strong>the</strong> toxicity <strong>of</strong> PQ. Most authors agree that upon entry <strong>into</strong> <strong>the</strong> cell, PQ<br />

undergoes a process <strong>of</strong> alternate reduction and reoxidation steps known as redox cycling<br />

(Fig. 11): PQ is reduced enzymatically, mainly by NADPH-cytochrome P-450<br />

reductase (Clejan and Cederbaum, 1989), NADH:ubiquinone oxidoreductase (complex<br />

I) (Fukushima et al., 1993; Yamada and Fukushima, 1993), xanthine oxidase (XO)<br />

(Winterbourn, 1981; Kelner et al., 1988; Waintrub et al., 1990; Kitazawa et al., 1991)<br />

and nitric oxide synthase (NOS) (Day et al., 1999) to form <strong>the</strong> PQ •+ plus NADP + or<br />

NAD + . It is generally accepted that PQ uses cellular diaphorases, which are a class <strong>of</strong><br />

enzymes that transfer electrons from NAD(P)H to small molecules, such as PQ (Dicker<br />

and Cederbaum, 1991; Aziz et al., 1994; Liochev and Fridovich, 1994; Day et al.,<br />

1999). The PQ •+ is <strong>the</strong>n rapidly reoxidized (returning to its original form) in <strong>the</strong><br />

presence <strong>of</strong> O2 [lungs exhibit high alveolar O2 tension (PAO2)] with <strong>the</strong> subsequent<br />

generation <strong>of</strong> O2 .- (Bus et al., 1974; Dicker and Cederbaum, 1991).


__________________________________________________Part I - General Introduction<br />

Interstitial<br />

space<br />

Cytoplasm<br />

NAD(P) +<br />

NAD(P)H<br />

HMP<br />

A<br />

C<br />

H 3<br />

C<br />

H 3<br />

+<br />

N<br />

+<br />

N<br />

H2O2 +<br />

Redox-Cycle<br />

Fe 2+<br />

PQ .+<br />

2+<br />

PQ<br />

+<br />

NADP<br />

NADPH<br />

PQ 2+<br />

O 2<br />

PUS<br />

.<br />

Gred<br />

N<br />

CH 3<br />

+<br />

N CH3 GSH<br />

GPx<br />

PQ .+<br />

O 2 .-<br />

O 2<br />

O 2<br />

.-<br />

SOD<br />

H 2 O 2<br />

CAT<br />

NO .<br />

HWR<br />

FR<br />

GSSG H2O O2 + H2O Fe3+ OH- + HO .<br />

+<br />

Type I, II and Clara<br />

cells membrance<br />

TOXICITY<br />

ONOO -<br />

.<br />

HO<br />

.<br />

HO<br />

Fig. 11 - Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. A. Cellular<br />

diaphorases, SOD, Superoxide dismutase; CAT, Catalase; GPx, Glutathione Peroxidase;<br />

Gred, Glutathione Reductase; PQ 2+ , Paraquat; PQ •+ , Paraquat monocation free radical;<br />

HMP, Hexose monophosphate pathway, FR; Fenton reaction; HWR, Haber-Weiss<br />

Reaction, PUS; polyamine uptake system.<br />

35


Part I - General Introduction __________________________________________________<br />

36<br />

The reaction between PQ •+ and O2 is very fast, with a rate constant <strong>of</strong> 7.7 × 10 8 M -<br />

1 s -1 (Farrington et al., 1973). The redox potential <strong>of</strong> PQ (PQ 2+ /PQ •+ ) is indeed very high<br />

(E0 ’ = −0.45 V), while that <strong>of</strong> molecular O2 (O2/O2 .- ) is lower (E0 ’ = −0.16 V), thus<br />

facilitating electron flow from <strong>the</strong> reduced PQ to O2. Provided that <strong>the</strong>re is sufficient<br />

NADPH as an electron donor, and O2 as an electron acceptor, PQ will play a catalytic<br />

role in this redox cycling process, generating O2 .- at <strong>the</strong> expense <strong>of</strong> NADPH. This <strong>the</strong>n<br />

sets in <strong>the</strong> well-known cascade leading to <strong>the</strong> production <strong>of</strong> o<strong>the</strong>r ROS, mainly<br />

hydrogen peroxide (H2O2), by dismutation <strong>of</strong> O2 .- , and HO . with <strong>the</strong> consequent cellular<br />

deleterious effects (Smith, 1987). This mechanism <strong>of</strong> action is also responsible for <strong>the</strong><br />

phytotoxic property <strong>of</strong> PQ (Dodge, 1971). Hydroxyl radicals may be generated by <strong>the</strong><br />

reaction <strong>of</strong> Haber-Weiss (Fig. 11). This reaction is very slow but may be catalyzed by<br />

traces <strong>of</strong> transition metal ions or metal chelates (Fenton reaction) (Winterbourn, 1981;<br />

Richmond and Halliwell, 1982; Kohen and Chevion, 1985b; Kohen and Chevion,<br />

1985c; Kohen and Chevion, 1985a).<br />

4.2 Biochemical consequences <strong>of</strong> <strong>the</strong> redox cycling process<br />

Most authors agree that redox cycling <strong>of</strong> PQ is a prerequisite for its toxicity likely<br />

to result in changes <strong>of</strong> <strong>the</strong> oxidative status. However, <strong>the</strong> critical biochemical events in<br />

<strong>the</strong> toxic process are far from clear. It should be stressed that <strong>the</strong> several processes need<br />

not necessarily to be mutually exclusive; it is quite possible that development <strong>of</strong><br />

irreversible cell damage is <strong>the</strong> consequence <strong>of</strong> tvarious events occurring independently<br />

<strong>of</strong> each o<strong>the</strong>r.<br />

4.2.1 Oxidation <strong>of</strong> NADPH<br />

A decrease in <strong>the</strong> ratio NADPH/NADP + on PQ-exposed lung tissue has been<br />

observed both in vitro (Sullivan and Montgomery, 1986) and in vivo (Witschi et al.,<br />

1977; Keeling et al., 1982). Although this is likely to be due partly to <strong>the</strong> oxidation <strong>of</strong><br />

NADPH (an essential co-factor required for <strong>the</strong> maintenance <strong>of</strong> normal biochemical and<br />

physiological processes) in <strong>the</strong> reduction <strong>of</strong> PQ, NADPH is also used as a c<strong>of</strong>actor <strong>of</strong><br />

glutathione reductase (Gred) in <strong>the</strong> regeneration <strong>of</strong> oxidized glutathione (GSSG) back to


__________________________________________________Part I - General Introduction<br />

reduced glutathione (GSH). GSSG is formed during <strong>the</strong> reduction <strong>of</strong> peroxides to<br />

alcohol, or during <strong>the</strong> detoxification <strong>of</strong> H2O2 <strong>into</strong> H2O by glutathione peroxidase (GPx).<br />

Several authors have observed a marked stimulation <strong>of</strong> <strong>the</strong> hexose monophosphate<br />

pathway (HMP) upon PQ treatment (Rose et al., 1976b; Bassett and Fisher, 1978;<br />

Keeling et al., 1982). Since this pathway represents <strong>the</strong> major cellular source <strong>of</strong><br />

NADPH, this probably reflects an effort <strong>of</strong> <strong>the</strong> lung to maintain levels <strong>of</strong> reducing<br />

equivalents under conditions <strong>of</strong> oxidative stress, by stimulation <strong>of</strong> glucose-6-phosphate<br />

dehydrogenase [(G6PD) <strong>the</strong> rate-limiting enzyme in <strong>the</strong> pathway)]. The studies <strong>of</strong><br />

Keeling and co-workers (Keeling and Smith, 1982) demonstrated a loss <strong>of</strong> NADPH in<br />

PQ-treated lungs within a few hours after PQ-exposure and before changes to <strong>the</strong><br />

alveolar epi<strong>the</strong>lium <strong>of</strong> <strong>the</strong> lung could be observed by electron microscopy. It has also<br />

been suggested that <strong>the</strong> activity <strong>of</strong> <strong>the</strong> enzyme G6PD is stimulated by GSSG, possibly<br />

through <strong>the</strong> formation <strong>of</strong> a mixed disulfide (Eggleston and Krebs, 1974). Since <strong>the</strong><br />

lowering <strong>of</strong> <strong>the</strong> NADPH/NADP + ratio is maintained despite <strong>the</strong> stimulation <strong>of</strong> <strong>the</strong> HMP,<br />

it is clear that this response is insufficient to overcome <strong>the</strong> oxidative stress. As<br />

suggested by Smith and Nemery (Smith and Nemery, 1986), it is perhaps ironic that<br />

stimulation <strong>of</strong> <strong>the</strong> HMP may, in fact, merely make available more NADPH for <strong>the</strong><br />

continued redox cycling <strong>of</strong> PQ and consequent ROS production. Assuming availability<br />

<strong>of</strong> NADPH and O2, <strong>the</strong> redox cycling <strong>of</strong> PQ continues on and on, with <strong>the</strong> continued<br />

depletion <strong>of</strong> NADPH, and generation <strong>of</strong> O2 .- .<br />

4.2.2 Oxidation <strong>of</strong> cellular thiol (SH) groups<br />

Several reports suggest that <strong>the</strong> onset <strong>of</strong> PQ toxicity is accompanied by a decrease<br />

in <strong>the</strong> levels <strong>of</strong> intracellular SH groups, predominantly through <strong>the</strong> oxidation <strong>of</strong> reduced<br />

GSH to GSSG and to <strong>the</strong> formation <strong>of</strong> protein mixed disulfides (Keeling and Smith,<br />

1982; Keeling et al., 1982). The oxidation <strong>of</strong> GSH to GSSG may occur through a direct<br />

effect <strong>of</strong> oxidizing species on <strong>the</strong> SH group. However, findings in GPx-deficient rat<br />

lungs (Glass et al., 1985) suggest that GSH is oxidized primarily through its role as a<br />

substrate in <strong>the</strong> GPx-mediated reduction <strong>of</strong> cellular H2O2. Both <strong>the</strong>ories suggest that <strong>the</strong><br />

prevention <strong>of</strong> <strong>the</strong> reduction <strong>of</strong> GSSG, formed as a consequence <strong>of</strong> redox cycling <strong>of</strong> PQ,<br />

results in enhanced toxicity. The mechanism underlying this phenomenon is unclear.<br />

One possibility is that <strong>the</strong> effect is due to depletion <strong>of</strong> GSH as <strong>the</strong> free SH group, thus<br />

37


Part I - General Introduction __________________________________________________<br />

preventing its participation in direct scavenging <strong>of</strong> free radicals and/or preventing<br />

removal <strong>of</strong> peroxides by GPx. A second possibility is that it is not <strong>the</strong> decreased<br />

availability <strong>of</strong> GSH, but <strong>the</strong> increase in GSSG levels that contributes to <strong>the</strong> toxic effect.<br />

Studies by Brigelius et al. (Brigelius et al., 1982) have shown that increases in cellular<br />

levels <strong>of</strong> GSSG lead to <strong>the</strong> formation <strong>of</strong> protein-glutathione mixed disulfides, possibly<br />

through <strong>the</strong> mediation <strong>of</strong> SH transferase enzymes. The structure and consequent<br />

activities <strong>of</strong> many cellular enzymes appear to be sensitive to mixed disulfide formation,<br />

some being inhibited while o<strong>the</strong>rs are stimulated as a consequence. Increased levels <strong>of</strong><br />

protein mixed disulfides have been observed in perfused liver (Brigelius et al., 1982)<br />

and in <strong>the</strong> lung (Keeling et al., 1982) <strong>of</strong> rats after exposure to PQ. Indeed, in <strong>the</strong> latter<br />

case, by administering various PQ doses, <strong>the</strong> authors were able to demonstrate a direct<br />

linear relationship between <strong>the</strong> increase in levels <strong>of</strong> mixed disulfides and stimulation <strong>of</strong><br />

<strong>the</strong> HMP activity. A similar relationship was demonstrated between mixed disulfide<br />

formation and inhibition <strong>of</strong> fatty acid syn<strong>the</strong>sis. This provides good evidence that<br />

oxidative changes occurring subsequently to PQ exposure result in cellular metabolism<br />

alterations.<br />

4.2.3 Oxidative damage to lipids, proteins and DNA<br />

Free radical-mediated membrane damage has been pointed by many authors as a<br />

critical event in <strong>the</strong> mechanism <strong>of</strong> PQ toxicity. According to this hypo<strong>the</strong>sis,<br />

electrophilic free radicals derived from <strong>the</strong> redox cycling <strong>of</strong> PQ are able <strong>of</strong> abstracting<br />

allylic hydrogen atoms from membrane-associated polyunsaturated fatty acids. In this<br />

manner, when <strong>the</strong> generation <strong>of</strong> radicals spreads, it results in alterations <strong>of</strong> membrane<br />

structure and ultimately, lipid peroxidation (LPO) (Yasaka et al., 1986). Indeed, HO .<br />

has been implicated in <strong>the</strong> initiation <strong>of</strong> membrane damage by LPO during <strong>the</strong> exposure<br />

to PQ in vitro (Bus et al., 1974; Bus et al., 1975; Shu et al., 1979) and in vivo (Bus et<br />

al., 1976; Burk et al., 1980; Dicker and Cederbaum, 1991). Curiously, clinical data<br />

concerning <strong>the</strong> LPO process in human PQ poisonings have been reported only rarely.<br />

Yasaka et al. (Yasaka et al., 1981; Yasaka et al., 1986) noted an increase in serum<br />

concentrations <strong>of</strong> malondialdehyde (MDA), a marker for LPO, in one case. Kurisaki<br />

(Kurisaki, 1985) reported an increase <strong>of</strong> MDA in <strong>the</strong> lung and liver in seven patients<br />

who died from acute PQ poisoning. Recently, Ranjbar et al. (Ranjbar et al., 2002)<br />

38


__________________________________________________Part I - General Introduction<br />

investigated <strong>the</strong> oxidative stress in blood samples <strong>of</strong> workers in a pesticide factory,<br />

formulating PQ products for use in agriculture. Controls were age-matched workers<br />

with no history <strong>of</strong> pesticide exposure. It was concluded that PQ-formulating factory<br />

workers have elevated LPO and decreased antioxidant power and total thiol (SH)<br />

groups in blood, revealing <strong>the</strong>ir liability to oxidative stress upon low but sustained<br />

exposure to PQ.<br />

The detection <strong>of</strong> hydrocarbons such as ethane or pentane in exhaled breath has<br />

attracted particular interest because <strong>the</strong>se volatile hydrocarbons are known to appear<br />

within seconds after <strong>the</strong> release <strong>of</strong> free radicals from tissues and reflect <strong>the</strong> extent <strong>of</strong><br />

peroxidized unsaturated fatty acids (Phillips, 1992; Kneepkens et al., 1994). Kazui et al.<br />

(Kazui et al., 1992) showed that <strong>the</strong> ethane in <strong>the</strong> expired breath (exEth) <strong>of</strong> rats reflects<br />

in vivo LPO. However, in ano<strong>the</strong>r study, and in spite <strong>of</strong> gross pulmonary damage<br />

revealed by <strong>the</strong> autopsy, following intratracheal rats exposure to PQ, exEth were not<br />

different from control animals (Schweich et al., 1994). The authors concluded that o<strong>the</strong>r<br />

markers than ethane must also be considered to detect this process in <strong>the</strong> lungs. Hong et<br />

al. (Hong et al., 2005) reported <strong>the</strong> first clinical trial attempt to evaluate <strong>the</strong> exEth as a<br />

clinical marker <strong>of</strong> <strong>the</strong> degree <strong>of</strong> lung damage following acute PQ poisoning in 21<br />

patients. The results indicated that even though <strong>the</strong> level <strong>of</strong> exEth was higher in <strong>the</strong><br />

nonsurvivor group than in <strong>the</strong> survivor group, it is nei<strong>the</strong>r an independent predictor <strong>of</strong><br />

survival nor a specific marker <strong>of</strong> lung injury in patients with acute PQ poisoning when it<br />

is measured 24 h after acute PQ poisoning. Ishii et al. (Ishii et al., 2002) collected lung,<br />

kidney, and liver at autopsy, from seven victims poisoned with PQ. The authors<br />

identified and reported an increase <strong>of</strong> oxysterols [detected as 7-ketocholesterol (7-keto)<br />

and 7-hydroxycholesterol (7α-OH and 7β-OH)] in <strong>the</strong> lung and kidney in response to<br />

PQ ingestion. These authors suggested that oxysterols are suitable lipid markers <strong>of</strong><br />

oxidative stress in man. Diene-conjugated 18:2Δ9,11-linoleic acid <strong>of</strong> plasma<br />

phospholipid <strong>of</strong> four patients (Situnayake et al., 1987) was also used as marker <strong>of</strong> LPO<br />

during <strong>the</strong> first few hours after <strong>the</strong> PQ poisoning.<br />

Besides lipids, ROS are also known to oxidatively modify DNA, carbohydrates<br />

and proteins. One such modification is <strong>the</strong> addition <strong>of</strong> carbonyl groups to amino acid<br />

residues in proteins. Free radical damage to proteins has been implicated in <strong>the</strong><br />

oxidative inactivation <strong>of</strong> several key metabolic enzymes. Fragmentation <strong>of</strong> polypeptide<br />

chains, increased sensitivity to denaturation, formation <strong>of</strong> protein–protein cross-linkages<br />

39


Part I - General Introduction __________________________________________________<br />

as well as modification <strong>of</strong> amino acids side chains to hydroxyl or carbonyl derivatives<br />

are possible outcomes <strong>of</strong> oxidation reactions (Dean et al., 1997).<br />

40<br />

Concerning DNA damage, PQ gave consistently positive results in assays for<br />

chromosomal damage (sister chromatid exchange, unscheduled DNA syn<strong>the</strong>sis and <strong>the</strong><br />

comet assay) in mammalian cells (S<strong>of</strong>uni et al., 1988; Ali et al., 1996; Dusinska et al.,<br />

1998). Using a human lung epi<strong>the</strong>lial-like cell line (L132), Takeyama et al. (Takeyama<br />

et al., 2004) showed that PQ-<strong>induced</strong> DNA damage by G1 arrest. The same study also<br />

demonstrated that PQ could induce single-stranded DNA breaks after 2 hours <strong>of</strong><br />

treatment. Tokunaga et al. (Tokunaga et al., 1997) studied <strong>the</strong> effect <strong>of</strong> PQ on base<br />

modifications, and showed an increase <strong>of</strong> 8-hydroxy-deoxyguanosine (8-OH-dG)<br />

formation in various rat <strong>organ</strong>s, particularly in brain, lung and heart. In contrast, <strong>the</strong><br />

formation <strong>of</strong> 8-hydroxy-guanosine (8-OH-G), a marker for <strong>the</strong> oxidative damage to<br />

RNA, was not significantly affected by PQ. These results indicate that PQ causes base<br />

modifications as well as strand breaks as a consequence <strong>of</strong> <strong>the</strong> oxidative damage to<br />

DNA. When PQ was incubated with lung homogenates prepared from mice in <strong>the</strong><br />

presence <strong>of</strong> calf thymus DNA, it caused damage to DNA in a concentration-dependent<br />

manner (Yamamoto and Mohanan, 2001). These results also suggest that HO . <strong>induced</strong><br />

by PQ probably account for <strong>the</strong> DNA damage, since damage was attenuated by <strong>the</strong> co-<br />

treatment with melatonin, a potent HO . scavenger.<br />

5. LUNG PATHOPHYSIOLOGY<br />

The mechanism <strong>of</strong> PQ toxicity is very similar to that <strong>of</strong> DQ at <strong>the</strong> molecular level.<br />

However, <strong>the</strong> critical target <strong>organ</strong> differs between <strong>the</strong> two compounds, so that <strong>the</strong><br />

mammalian toxicology is quite different. While both herbicides affect <strong>the</strong> kidneys, PQ<br />

is selectively accumulated in <strong>the</strong> lungs through a saturable uptake process (Rose et al.,<br />

1974; Rose et al., 1976a; Smith, 1982; Smith and Nemery, 1992) and <strong>the</strong> systemic<br />

toxicity <strong>of</strong> PQ is dominated by lung toxicity. The pathological changes in <strong>the</strong> lung<br />

provoked by PQ have been investigated in various species <strong>of</strong> experimental animals. The<br />

rat, mouse, dog, and monkey, develop lung damage similar to that observed in man<br />

(Conning et al., 1969; Murray and Gibson, 1972). The pathogenesis <strong>of</strong> PQ toxicity has<br />

been most extensively studied in <strong>the</strong> rat. There are two distinct phases in <strong>the</strong>


__________________________________________________Part I - General Introduction<br />

development <strong>of</strong> pulmonary lesions (Smith et al., 1974b; Smith and Heath, 1976). These<br />

coincide with <strong>the</strong> early and late clinical stages. The initial stage involves acute damage<br />

to several <strong>organ</strong>s, including liver, heart, kidneys, and lungs. Depending on <strong>the</strong> amount<br />

<strong>of</strong> PQ ingested, death may occur during this period and is associated with pulmonary,<br />

renal, and circulatory failure (Smith and Heath, 1976) (Table 5). Patients surviving this<br />

stage generally show a period <strong>of</strong> improvement. However, in most cases this is merely<br />

<strong>the</strong> prelude to <strong>the</strong> onset <strong>of</strong> <strong>the</strong> second stage, which involves damage almost exclusively<br />

to <strong>the</strong> lungs. Extensive pulmonary fibrosis ensues, resulting in dyspnea, cyanosis, and<br />

eventually death from respiratory failure. Neverthless, it has been found that some<br />

species do not develop lung lesions. For example, <strong>the</strong> rabbit lung (Butler and<br />

Kleinerman, 1971) was not damaged by a single dose <strong>of</strong> PQ, although chronic administration<br />

to rabbits can induce lung damage (Seidenfeld et al., 1978).<br />

Table 5 – Phases <strong>of</strong> <strong>paraquat</strong> toxicity and associated clinical effects. GIT,<br />

gastrointestinal tract; CNS, central nervous system; 1 doses as low as 4 mg/Kg can cause<br />

death (Driesbach, 1983).<br />

Phases <strong>of</strong><br />

Toxicity<br />

I.<br />

Asymptomatic<br />

or mild<br />

II. Moderate to<br />

severe<br />

Ingested<br />

PQ ion<br />

dose<br />

(mg/Kg<br />

b.w.)<br />

20-30<br />

but


Part I - General Introduction __________________________________________________<br />

III. Severe:<br />

acute<br />

fulminant<br />

toxicity<br />

42<br />

>40-55<br />

5.1 Destructive phase<br />

>15 mL <strong>of</strong><br />

20% (m/v)<br />

concentrate<br />

Nausea, emesis and<br />

diarrhea are followed by<br />

multi<strong>organ</strong>ic failure<br />

(hepatic, renal, adrenal,<br />

pancreatic, CNS,<br />

cardiac and respiratory<br />

failure). Patients do not<br />

survive long enough to<br />

demonstrate pulmonary<br />

fibrosis.<br />

Marked<br />

ulcerations as<br />

in phase II.<br />

Esophageal<br />

perforation<br />

and<br />

mediastinitis<br />

can occur<br />

within 2–3<br />

days <strong>of</strong> <strong>the</strong><br />

ingestion<br />

Death<br />

usually<br />

occurs<br />

within 24<br />

hours<br />

(generally<br />

not<br />

delayed<br />

for more<br />

than a few<br />

days)<br />

The first toxicological effects to <strong>the</strong> lung correspond to a destructive phase in<br />

which <strong>the</strong> alveolar type I and type II epi<strong>the</strong>lial cells are destroyed. This occurs within 1-<br />

3 days <strong>of</strong> dosing, although <strong>the</strong> speed at which it occurs depends on <strong>the</strong> given dose and<br />

<strong>the</strong> route <strong>of</strong> administration. Irrespective <strong>of</strong> <strong>the</strong>se factors, <strong>the</strong> earliest observed<br />

pulmonary changes caused by PQ occurs in <strong>the</strong> type I alveolar epi<strong>the</strong>lial cells, which<br />

exhibit swelling (Kimbrough and Gaines, 1970; Sykes et al., 1977) accompanied by<br />

increases in <strong>the</strong>ir content <strong>of</strong> mitochondria and ribosomes, changes suggestive <strong>of</strong><br />

increased metabolic activity (Smith et al., 1974b). Cell damage initially appears as<br />

mitochondrial swelling, followed by overt cell degeneration and cytoplasmic edema.<br />

The latter, results in bulging <strong>of</strong> <strong>the</strong> cytoplasm <strong>into</strong> <strong>the</strong> alveolar space, and progresses to<br />

<strong>the</strong> rupture <strong>of</strong> <strong>the</strong> type I cell to expose <strong>the</strong> basement membrane (Smith and Heath,<br />

1976). Early damage to type I alveolar cells by PQ may be explained by <strong>the</strong> fact that<br />

<strong>the</strong>y cover a large surface area (approximately 93% <strong>of</strong> <strong>the</strong> alveolar epi<strong>the</strong>lial surface<br />

area), representing 33% <strong>of</strong> alveolar epi<strong>the</strong>lial cells. The main function <strong>of</strong> <strong>the</strong> type I<br />

alveolar cells, which are flat and actually form <strong>the</strong> alveolar vesicle is <strong>the</strong> gas exchange<br />

between <strong>the</strong> air space and <strong>the</strong> capillaries. PQ deeply compromises lung function since<br />

<strong>the</strong> beginning <strong>of</strong> its toxic effects. The alveolar type II cell represents <strong>the</strong> only o<strong>the</strong>r lung<br />

cell type to show overt damage during this early phase <strong>of</strong> PQ toxicity. Damage to <strong>the</strong><br />

type II cell appears to lag slightly behind <strong>the</strong> type I cell injury, and first involves<br />

mitochondrial swelling and loss <strong>of</strong> <strong>the</strong> contents <strong>of</strong> <strong>the</strong> characteristic lamellar bodies<br />

(which are believed to contain surfactant) before frank cell destruction (Smith and<br />

Heath, 1976). The type II cells are more round shaped and located at <strong>the</strong> distal border <strong>of</strong>


__________________________________________________Part I - General Introduction<br />

<strong>the</strong> alveolar vesicles. They account for <strong>the</strong> remaining 7% by surface area and 67% by<br />

epi<strong>the</strong>lial cell number. Their main functions are surfactant secretion, active transport <strong>of</strong><br />

water and ions, and epi<strong>the</strong>lial regeneration. The role <strong>of</strong> <strong>the</strong> surfactant (phospholipids,<br />

mainly phosphatidylcholine) is to form a thin film on top <strong>of</strong> a thin aqueous layer that<br />

covers <strong>the</strong> epi<strong>the</strong>lial cells. This decreases <strong>the</strong> surface tension and thus prevents <strong>the</strong> lung<br />

collapse during expiration. They also act as a defense against toxic agents in<br />

consequence <strong>of</strong> <strong>the</strong>ir particularly richness in NADPH-cytochrome P-450 reductase, and<br />

may undergo mitotic division and replace type I damaged cells. Notwithstanding some<br />

authors have observed morphological changes, including swelling (Brooks, 1971;<br />

Fukuda et al., 1985) and even vacuolization (Modee et al., 1972) <strong>of</strong> <strong>the</strong> capillary<br />

endo<strong>the</strong>lium, <strong>the</strong> weight <strong>of</strong> <strong>the</strong> evidence suggests that <strong>the</strong>se cells initially remain<br />

essentially undamaged, even at an ultrastructural level (Vijeyaratnam and Corrin, 1971;<br />

Sykes et al., 1977). Certainly, <strong>the</strong> overt damage and destruction seen early in <strong>the</strong><br />

epi<strong>the</strong>lium do not manifest <strong>the</strong>mselves in <strong>the</strong> endo<strong>the</strong>lium. Dearden et al. (Dearden et<br />

al., 1982) observed endo<strong>the</strong>lial damage in rats only 48 hours after i.p. administration <strong>of</strong><br />

PQ. In endo<strong>the</strong>lial cells, on <strong>the</strong> septal side <strong>of</strong> <strong>the</strong> capillaries, <strong>the</strong> number <strong>of</strong> pinocytotic<br />

vesicles significantly increased from 48 to 96 hours post-PQ. In endo<strong>the</strong>lium adjacent to<br />

damaged epi<strong>the</strong>lium, abnormalities included hydration, fragmentation, discontinuity,<br />

and widened intercellular junctions; <strong>the</strong>se were maximal 72-96 hours post-PQ. These<br />

and o<strong>the</strong>r authors concluded that although o<strong>the</strong>r <strong>mechanisms</strong> are probably important,<br />

damaged pulmonary capillary endo<strong>the</strong>lium seems to be a factor favoring <strong>the</strong> onset <strong>of</strong> an<br />

alveolitis, which is characterized by <strong>the</strong> production <strong>of</strong> a pulmonary hemorrhage<br />

proteinaceous edema and by <strong>the</strong> infiltration <strong>of</strong> <strong>the</strong> interstitial tissue and air spaces <strong>of</strong> <strong>the</strong><br />

lung with inflammatory cells (Vijeyaratnam and Corrin, 1971; Sykes et al., 1977).<br />

However, it should be noted that endo<strong>the</strong>lial cell damage is notoriously difficult to<br />

demonstrate morphologically, even in instances in which <strong>the</strong>re is functional evidence <strong>of</strong><br />

microvascular impairment (Pietra, 1984). It has also been suggested that <strong>the</strong> destruction<br />

<strong>of</strong> <strong>the</strong> surfactant-producing type II cells results in increased surface tension within <strong>the</strong><br />

alveoli, and that this draws fluid from <strong>the</strong> capillaries to produce edema (Gardiner,<br />

1972). Alternatively, edema may also result from permeability changes in <strong>the</strong> alveolar<br />

wall subsequent to type I cell damage (Sykes et al., 1977). The inflammatory response<br />

that arises during this destructive phase, which is maintained throughout <strong>the</strong><br />

proliferative phase, involves a rapid and extensive influx <strong>of</strong> inflammatory cells, mainly<br />

<strong>of</strong> polymorphonuclear leukocytes, macrophages (Clark et al., 1966; Brooks, 1971;<br />

43


Part I - General Introduction __________________________________________________<br />

Smith and Heath, 1974; Smith et al., 1974b; Sykes et al., 1977; Wasserman and Block,<br />

1978; Fukuda et al., 1985), and eosinophils (Clark et al., 1966; Vijeyaratnam and<br />

Corrin, 1971; Gardiner, 1972; Modee et al., 1972; Pietra, 1984; Fukuda et al., 1985;<br />

Candan and Alagozlu, 2001), <strong>into</strong> <strong>the</strong> interstitium and alveolar spaces. Most rats die<br />

within a few days after PQ exposure as a consequence <strong>of</strong> this extensive alveolitis and<br />

pulmonary edema. In human <strong>into</strong>xication cases, <strong>the</strong> edema is generally not as extensive<br />

as seen in <strong>the</strong> rodent lung and when it develops it is usually subject to clinical<br />

management.<br />

5.2 Proliferative phase<br />

The second phase <strong>of</strong> PQ-<strong>induced</strong> lung toxicity involves <strong>the</strong> development <strong>of</strong> an<br />

extensive fibrosis in <strong>the</strong> lung, which is probably a compensatory repair mechanism to<br />

<strong>the</strong> damaged alveolar epi<strong>the</strong>lial cells during alveolitis (Smith and Heath, 1976). If <strong>the</strong><br />

degree <strong>of</strong> lung exposure to PQ is high, <strong>the</strong> alveolitis will be more widespread and<br />

severe, <strong>the</strong>reby resulting in a more extensive fibrosis and severe anoxia. Thus, <strong>the</strong><br />

fibrosis may be part <strong>of</strong> <strong>the</strong> normal reparative response <strong>of</strong> <strong>the</strong> lung to severe and<br />

extensive damage. The fibrosis associated with PQ toxicity is not exceptional or<br />

peculiar to <strong>the</strong> effects <strong>of</strong> PQ but is a response to an acute alveolitis that can also be<br />

<strong>induced</strong> by many o<strong>the</strong>r pulmonary toxins. The onset <strong>of</strong> <strong>the</strong> proliferative phase occurs<br />

several days after PQ ingestion. The earliest morphological indication <strong>of</strong> fibrotic<br />

development is <strong>the</strong> appearance <strong>of</strong> many pr<strong>of</strong>ibroblasts in <strong>the</strong> alveolar spaces (Smith and<br />

Heath, 1976). These cells undergo rapid proliferation and differentiation to mature<br />

fibroblasts, which lay down collagen and ground substance to produce fibrosis. This<br />

fibrotic proliferation is very rapid, resulting in <strong>the</strong> loss <strong>of</strong> <strong>the</strong> normal alveolar<br />

architecture, interfering with gaseous exchange, and subsequently causing death from<br />

anoxia. Smith and Heath (Smith and Heath, 1976) claimed that <strong>the</strong> localization <strong>of</strong> <strong>the</strong><br />

fibroblasts (both immature and mature) and <strong>of</strong> <strong>the</strong> subsequent fibrotic lesion is entirely<br />

intra-alveolar. O<strong>the</strong>r <strong>research</strong>ers have described interstitial in addition to intra-alveolar<br />

fibrosis (Fukuda et al., 1985), although <strong>the</strong>y suggest that <strong>the</strong> intra-alveolar component is<br />

never<strong>the</strong>less more deleterious, because it is <strong>the</strong> latter that results in obliteration <strong>of</strong> <strong>the</strong><br />

alveoli. The <strong>mechanisms</strong> for <strong>the</strong> development <strong>of</strong> this obliterating fibrosis are still poorly<br />

understood. Predominantly, in <strong>the</strong> event <strong>of</strong> interstitial or intra-alveolar fibrosis,<br />

44


__________________________________________________Part I - General Introduction<br />

whatever <strong>the</strong> cause, <strong>the</strong> normal architecture <strong>of</strong> <strong>the</strong> lung is destroyed due to <strong>the</strong><br />

proliferation <strong>of</strong> fibroblasts and deposition <strong>of</strong> collagen, <strong>the</strong>reby reducing <strong>the</strong><br />

effectiveness <strong>of</strong> gaseous exchange, leading to death as a consequence <strong>of</strong> severe anoxia.<br />

Using o<strong>the</strong>r experimental systems, not involving PQ, Witschi and co-workers have<br />

proposed that pulmonary fibrosis occurs when re-epi<strong>the</strong>lialisation subsequent to<br />

epi<strong>the</strong>lial damage is compromised in some manner (Witschi et al., 1980). Such a<br />

process would certainly appear to hold true also in <strong>the</strong> case <strong>of</strong> PQ, since <strong>the</strong> replacement<br />

<strong>of</strong> damaged type I cells (which constitute <strong>the</strong> majority <strong>of</strong> <strong>the</strong> epi<strong>the</strong>lial surface area) is<br />

prevented by destruction <strong>of</strong> <strong>the</strong>ir progenitor type II cells. Moreover, Fukuda et al.<br />

(Fukuda et al., 1985) suggested that secretion <strong>of</strong> proteolytic enzymes by stimulated<br />

inflammatory cells may result in degradation <strong>of</strong> alveolar basement membranes denuded<br />

by loss <strong>of</strong> <strong>the</strong> epi<strong>the</strong>lium, and that this may also inhibit epi<strong>the</strong>lial regeneration. This is<br />

consistent with <strong>the</strong> idea that within a given area <strong>of</strong> damage, re-epi<strong>the</strong>lialisation and<br />

fibrotic proliferation represent mutually exclusive endpoints, and that <strong>the</strong> balance<br />

between <strong>the</strong> two is governed by <strong>the</strong> degree <strong>of</strong> epi<strong>the</strong>lial damage. Thus, if re-<br />

epi<strong>the</strong>lialisation is delayed (due to type II cell damage or to destruction <strong>of</strong> <strong>the</strong> basement<br />

membrane), fibrosis may occur. However, it appears that at least in <strong>the</strong> case <strong>of</strong> PQ,<br />

o<strong>the</strong>r factors may also play a role. In <strong>the</strong>ir review, Smith and Heath (Smith and Heath,<br />

1976) concluded that development <strong>of</strong> PQ-<strong>induced</strong> pulmonary fibrosis is independent <strong>of</strong><br />

alveolar damage. They suggested that PQ may itself initiate <strong>the</strong> influx <strong>of</strong> pro-fibroblasts<br />

(Smith and Heath, 1976). This was evidenced to some degree by Conning et al.<br />

(Conning et al., 1969), who demonstrated that macrophages treated with PQ caused a<br />

more rapid proliferation <strong>of</strong> cultured fibroblasts than did untreated macrophages, as well<br />

as by Schoenberger et al. (Schoenberger et al., 1984), who showed <strong>the</strong> release <strong>of</strong> a<br />

fibroblast growth factor by PQ-exposed macrophages. Despite some advances toward<br />

understanding <strong>the</strong> nature <strong>of</strong> <strong>the</strong> PQ-<strong>induced</strong> fibrotic lung lesion, <strong>the</strong> ultimate<br />

<strong>mechanisms</strong> underlying this process, as with pulmonary fibrosis in general, remain<br />

elusive (Gharaee-Kermani and Phan, 2005).<br />

45


Part I - General Introduction __________________________________________________<br />

46<br />

6. OBSERVATIONS IN ANIMALS AND HUMANS<br />

Several studies have been performed to evaluate <strong>the</strong> acute toxicity <strong>of</strong> PQ<br />

administered by a variety <strong>of</strong> routes. An overall picture <strong>of</strong> <strong>the</strong> obtained results is<br />

summarized in Table 6. Both PQ sulphate and dichloride salts are equally toxic when<br />

expressed on <strong>the</strong> basis <strong>of</strong> PQ ion (Clark et al., 1966). There is a great variation in LD50<br />

values, depending upon <strong>the</strong> investigator, <strong>the</strong> laboratory where <strong>the</strong> work was done and as<br />

result <strong>of</strong> <strong>the</strong> inherent differences in sensitivity between species, route <strong>of</strong> administration,<br />

and reproductive state. Evidence also exists <strong>of</strong> young animals being more susceptible<br />

(Clark et al., 1966). Also, individual animals <strong>of</strong> <strong>the</strong> same species show an unusually<br />

large variation in <strong>the</strong> time from dosing to death following identical dosage. When rats<br />

were weighed daily following a single oral or i.v. dose at a rate that would kill only a<br />

fraction <strong>of</strong> <strong>the</strong> tested animals, it was found that those minimally affected lost weight<br />

only briefly, and <strong>the</strong>n began gradually to regain, whereas those that were severely<br />

affected continued to lose weight (Sharp et al., 1972). Rats lost at least as much BW as<br />

would be expected during total deprivation <strong>of</strong> food and water (Peters, 1967). Weights <strong>of</strong><br />

<strong>the</strong> two groups were statistically different after <strong>the</strong> first day following oral<br />

administration and on all days following i.v. administration. Weight loss appeared to be<br />

<strong>the</strong> result <strong>of</strong> lower food intake. Whe<strong>the</strong>r as <strong>the</strong> result <strong>of</strong> differences in absorption<br />

following oral administration or <strong>of</strong> greater excretion or sequestration regardless <strong>of</strong> <strong>the</strong><br />

route <strong>of</strong> administration, minimally affected rats contained less PQ in <strong>the</strong>ir lungs, kidneys,<br />

and stomach during days 1-8 than did severely affected ones, including those that<br />

died (Sharp et al., 1972). In rats, this interval varies from 2 to 12 days, with some<br />

tendency for <strong>the</strong> deaths to be concentrated in an early and late peak (Clark et al., 1966).<br />

After rats had inhaled PQ, clinical signs and post-mortem markers <strong>of</strong> toxicity were<br />

similar to those seen after oral, s.c. or i.p. administration. Aerosol LC50 values in PQ<br />

toxicity tests with mammals were directly related to <strong>the</strong> duration <strong>of</strong> exposure, PQ<br />

concentration in spray, and particle size [3 µm (diameter) seemed most effective<br />

(Haley, 1979)].


__________________________________________________Part I - General Introduction<br />

Table 6 – Paraquat LD50 in various species. NS, not stated; M, male; F, female; a dose<br />

quoted as <strong>paraquat</strong> ion; b as dimethylsulphate.<br />

Species Strain Sex Route<br />

Mouse NS NS<br />

Rat<br />

per os 120<br />

i.p. 30<br />

i.v. 180<br />

LD50 (mg/Kg<br />

b.w.) (95%<br />

confidence<br />

interval)<br />

Swiss-Webster M i.p. 39 (32.5-46.8)<br />

Swiss-Webster F i.p. 30 (26.3-34.2)<br />

NS F i.p. 19 (16-21) a<br />

NS F i.p. 16 (10-26)<br />

NS NS i.v. 21<br />

NS F per os 112 (104-122) a<br />

NS F per os 150 (139-162) a<br />

Sherman M per os 100 b<br />

Sherman F per os 110 b<br />

NS F per os 150(110-173)<br />

Sprague-Dawley M per os 126<br />

NS NS per os 57<br />

Sherman M dermal 80 b<br />

Sherman F dermal 90 b<br />

Rabbit NS M per os 50 (45-58)<br />

Reference<br />

(Orme and<br />

Kegley)<br />

(Sinow and<br />

Wei, 1973)<br />

(Bus et al.,<br />

1976)<br />

(Clark et al.,<br />

1966)<br />

(Mehani,<br />

1972)<br />

(Orme and<br />

Kegley)<br />

(Clark et al.,<br />

1966)<br />

(Clark et al.,<br />

1966)<br />

(Kimbrough<br />

and Gaines,<br />

1970)<br />

(Kimbrough<br />

and Gaines,<br />

1970)<br />

(Mehani,<br />

1972)<br />

(Murray and<br />

Gibson,<br />

1972)<br />

(Orme and<br />

Kegley)<br />

(Kimbrough<br />

and Gaines,<br />

1970)<br />

(Kimbrough<br />

and Gaines,<br />

1970)<br />

(Mehani,<br />

1972)<br />

47


Part I - General Introduction __________________________________________________<br />

48<br />

NS M i.p. 25 (15-30)<br />

NS dermal 236<br />

Cats NS F per os 35 (27-46) a<br />

Dog<br />

Monkeys<br />

Guineapigs<br />

Beagles<br />

M s.c. 1.8 (1.0-6.1)<br />

F s.c. 3.5 (2.4-10.1)<br />

NS NS oral 25<br />

Cynomolgus<br />

(Macaca<br />

fascicularis)<br />

M and F per os 50<br />

M per os 70 a<br />

NS M per os 30 (22-41) a<br />

Sprague-Dawley M and F per os 22<br />

NS F i.p. 3 a<br />

(Mehani,<br />

1972)<br />

(Clark et al.,<br />

1966)<br />

(Clark et al.,<br />

1966)<br />

(Nagata et<br />

al., 1992)<br />

(Nagata et<br />

al., 1992)<br />

(Orme and<br />

Kegley)<br />

(Murray and<br />

Gibson,<br />

1972)<br />

(Purser and<br />

Rose, 1979)<br />

(Clark et al.,<br />

1966)<br />

(Murray and<br />

Gibson,<br />

1972)<br />

(Clark et al.,<br />

1966)<br />

Although histopathological alterations are generally similar among <strong>the</strong> rat, dog,<br />

monkey and mice (Clark et al., 1966; Murray and Gibson, 1972), Butler (Butler, 1975)<br />

found that <strong>the</strong> Syrian hamster is relatively resistant to interstitial fibrosis. Butler and<br />

Kleinerman (Butler and Kleinerman, 1971) also reported that rabbits did not develop <strong>the</strong><br />

pulmonary changes typical <strong>of</strong> PQ poisoning in o<strong>the</strong>r species, despite doses <strong>of</strong> 2–100<br />

mg/Kg BW being administered i.p. and sacrifice <strong>of</strong> animals being delayed up to 1<br />

month. The only findings in <strong>the</strong> lungs were occasional small interstitial infiltrates <strong>of</strong><br />

lymphocytes and plasma cells, minimal alveolar hyperplasia, and some alveolar<br />

macrophages.<br />

Human deaths from acute PQ poisoning started to be reported in <strong>the</strong> medical<br />

literature in 1966, when Bullivant (Bullivant, 1966) reported two fatalities in New<br />

Zealand due to accidental ingestion <strong>of</strong> PQ and mentioned a previous fatality that had<br />

occurred in Ireland in 1964. During 1967 and 1968, <strong>the</strong>re were no less than 13 cases <strong>of</strong>


__________________________________________________Part I - General Introduction<br />

PQ poisonings reported in <strong>the</strong> literature, 9 <strong>of</strong> which were fatal (Malone et al., 1971).<br />

Most <strong>of</strong> <strong>the</strong> initial reports on PQ involved accidental poisonings, usually related to<br />

storage <strong>of</strong> <strong>the</strong> herbicide in s<strong>of</strong>t drink, wine, beer or o<strong>the</strong>r common beverage bottles, or<br />

in inappropriately labelled containers as well as due to poor worker-protection practices.<br />

PQ soon gained a reputation, not only among <strong>the</strong> medical community but also <strong>the</strong><br />

general public, as being one <strong>of</strong> <strong>the</strong> most toxic substances available and by <strong>the</strong> apparent<br />

inability <strong>of</strong> <strong>the</strong>rapeutic efforts to alter <strong>the</strong> outcome.<br />

The LD50 <strong>of</strong> PQ is 3 to 5 g for human adults (Smith, 1988b). As little as a<br />

mouthful (approximately 20 mL) <strong>of</strong> a 20% solution <strong>of</strong> PQ produces a dose <strong>of</strong> ~55<br />

mg/Kg in an average 70-Kg adult, which may be fatal. The lowest fatal dose recorded<br />

for adult humans is 17 mg/Kg, but lower doses may be fatal for children (Wesseling et<br />

al., 2001). One tea spoon <strong>of</strong> PQ may kill a starved children (Harley et al., 1977). PQ<br />

ingestion is one <strong>of</strong> <strong>the</strong> leading methods <strong>of</strong> suicide in countries such as Taiwan, Japan,<br />

Malaysia, <strong>the</strong> West Indies, and Samoa. In Japan, 1,200 to 1,500 deaths/year from PQ<br />

ingestions were reported in <strong>the</strong> 1980s (Onyon and Volans, 1987). In a retrospective<br />

study on 639 pesticides analysis requests between January 2000 and December 2002 in<br />

<strong>the</strong> centre <strong>of</strong> Portugal, Teixeira et al. (Teixeira et al., 2004) reported 31 deaths due to<br />

PQ in 111 positive pesticide <strong>into</strong>xication cases.<br />

6.1 Clinical symptoms and manifestations <strong>of</strong> <strong>paraquat</strong> <strong>into</strong>xication<br />

The effects <strong>of</strong> PQ are local and systemic, <strong>the</strong> former being concentration<br />

dependent, while <strong>the</strong> latter are dose-dependent (Proudfoot, 1999). Although <strong>the</strong> local<br />

effects can be severe, it is <strong>the</strong> systemic effects, largely referable to <strong>the</strong> respiratory<br />

system, that are potentially lethal. Findings suggest that PQ may cause fatal poisonings<br />

by ingestion <strong>of</strong> small amounts, by dermal absorption <strong>of</strong> PQ, and possibly by inhalation<br />

(Wesseling et al., 1997). The clinical features <strong>of</strong> <strong>paraquat</strong> poisoning are summarized in<br />

Table 7.<br />

Table 7 – Clinical features <strong>of</strong> <strong>paraquat</strong> poisonings.<br />

Cardiovascular<br />

Hypovolemia, shock, dysrhythmias<br />

49


Part I - General Introduction __________________________________________________<br />

Central nervous<br />

Coma, convulsions, cerebral edema<br />

Dermatologic<br />

Corrosion <strong>of</strong> skin, nails, cornea, conjunctiva, and nasal mucosa<br />

Endocrine<br />

Adrenal insufficiency caused by adrenal necrosis as part <strong>of</strong> <strong>multiple</strong> <strong>organ</strong> failure<br />

Gastrointestinal<br />

Oropharyngeal ulceration and corrosion; nausea, vomiting, hematemesis, diarrhea,<br />

dysphagia, perforation <strong>of</strong> esophagus, pancreatitis, centrilobular hepatic necrosis,<br />

cholestasis<br />

Genitourinary<br />

Oliguric or nonoliguric renal failure caused by acute tubular necrosis; proximal<br />

tubular dysfunction<br />

Hematopoietic<br />

Leukocytosis early, anemia late<br />

Respiratory<br />

Cough, aphonia, prominent pharyngeal membranes (pseudodiph<strong>the</strong>ria), mediastinitis,<br />

pneumothorax, hemoptysis, acute lung injury, hemorrhage, pulmonary fibrosis<br />

6.1.1 Poisoning by <strong>the</strong> oral route<br />

Acute PQ poisonings are mostly due to ingestion <strong>of</strong> <strong>the</strong> concentrate liquid<br />

herbicide formulations, since granular formulations containing 2.5 or 5 g (w/w) could<br />

only be swallowed in quantity after being dissolved in water.<br />

The symptomatology <strong>of</strong> human PQ poisonings can be divided <strong>into</strong> three different<br />

presentations depending on <strong>the</strong> amount ingested (Table 5) (Vale et al., 1987; Pond,<br />

1990; Bismuth et al., 1995).<br />

6.1.1.1 Severe toxicity<br />

Patients who ingest greater than 40 mg/Kg (>15 mL <strong>of</strong> a 20% solution for a 70-Kg<br />

patient) usually die within hours to a few days, at most (Bismuth et al., 1982; Pond,<br />

1990). These patients experience <strong>multiple</strong> <strong>organ</strong> failure, including acute respiratory<br />

distress syndrome (ARDS), cerebral edema, myocardial necrosis, and cardiac,<br />

50


__________________________________________________Part I - General Introduction<br />

neurologic, adrenal, pancreatic, hepatic (with jaundice) and renal failure (Nagi, 1970;<br />

Russell et al., 1981; Bismuth et al., 1982; Reif and Lewinsohn, 1983; Pond, 1990;<br />

Florkowski et al., 1992). Death may occur even before <strong>the</strong> development <strong>of</strong> significant<br />

chest radiographic abnormalities (Pond, 1990). Alveolitis is observed, with clinical<br />

signs <strong>of</strong> acute noncardiogenic pulmonary edema and rapidly progressive hypoxemia,<br />

even in patients treated with salt and fluid restriction. Acute pneumonitis, shock,<br />

metabolic acidosis, and convulsions have been reported. Nausea, vomiting, and<br />

abdominal pain are also present. Bloody diarrhea may be present.<br />

6.1.1.2 Moderate toxicity – <strong>the</strong> typical <strong>into</strong>xication<br />

Patients who ingest >20-30 but


Part I - General Introduction __________________________________________________<br />

several cases <strong>of</strong> esophageal and gastric ulceration preceding perforation and massive<br />

GIT hemorrhage have been reported (Malone et al., 1971; Ackrill et al., 1978). O<strong>the</strong>r<br />

signs <strong>of</strong> GIT irritation such as nausea and vomiting may occur. Vomiting almost always<br />

ensues, even in <strong>the</strong> absence <strong>of</strong> emetic agents in <strong>the</strong> commercial preparation.<br />

Secondarily, abdominal colic and diarrhea are noted occasionally.<br />

6.1.1.2.2 Second phase<br />

Between <strong>the</strong> second and fifth days following ingestion, renal failure and<br />

hepatocellular necrosis develop. Functional renal insufficiency is <strong>of</strong>ten noted, caused<br />

partly by hypovolaemia secondary to GIT fluid losses and a decreased or total lack <strong>of</strong><br />

oral fluid intake. PQ itself has direct renal toxicity. It generally causes a pure<br />

tubulopathy with proximal predominance. Such renal tubulopathies usually evolve - as<br />

with all causes <strong>of</strong> tubular necrosis - to full recovery without sequelae. Although <strong>the</strong><br />

degree <strong>of</strong> renal failure may be mild by most standards, renal failure impairs <strong>the</strong> only<br />

route <strong>of</strong> excretion available and <strong>the</strong>refore may contribute significantly to <strong>the</strong> mortality<br />

produced by PQ. PQ poisoned patients frequently recover renal function by <strong>the</strong> time <strong>of</strong><br />

death. However, in <strong>the</strong>se fatal cases, renal tubular damage has been noted at autopsy.<br />

Although acute tubular necrosis is <strong>the</strong> most common form <strong>of</strong> <strong>the</strong> renal involvement,<br />

various pictures have also been observed, such as tubular dysfunction mimicking<br />

Fanconi syndrome (Vaziri et al., 1979; Stratta et al., 1988). In a study reporting <strong>the</strong><br />

nephrotoxicity <strong>of</strong> PQ in vitro and in vivo, proximal tubular function was monitored by<br />

measuring <strong>the</strong> accumulation <strong>of</strong> PAH and NMN using renal cortical slices from Swiss-<br />

Webster mice poisoned with PQ at <strong>the</strong> LD50 for i.p. administration (50 mg/Kg BW)<br />

(Ecker et al., 1975). Tubular function in intact Swiss-Webster mice was estimated using<br />

disappearance <strong>of</strong> phenolsulfthalein and [ 14 C]PQ from plasma in vivo. Glomerular<br />

function was estimated using disappearance <strong>of</strong> othalamate from <strong>the</strong> plasma <strong>of</strong> animal’s<br />

injected i.v. with PQ at a dose <strong>of</strong> 50 mg/Kg BW. Accumulation <strong>of</strong> PAH and NMN by<br />

renal cortical slices in vitro was not greatly altered. In vivo disappearance <strong>of</strong><br />

phenolsulfthalein and [ 14 C]PQ from plasma was greatly reduced, but iothalamate<br />

disappearance was little affected. These authors concluded that <strong>the</strong> nephrotoxicity<br />

attributable to PQ affects primarily <strong>the</strong> proximal tubule (Ecker et al., 1975). These<br />

findings are supported by in vitro experiments in which <strong>the</strong> proximal renal epi<strong>the</strong>lial<br />

52


__________________________________________________Part I - General Introduction<br />

cell line (LLC-PK1) was found to be more susceptible to <strong>the</strong> toxic effects <strong>of</strong> PQ when<br />

compared to a distal epi<strong>the</strong>lial cell line, MDCK (Chan et al., 1996a). It has been noted<br />

that <strong>the</strong> uptake <strong>of</strong> PQ by rat renal tubular cells in culture is saturable (Chan et al.,<br />

1996b). Besides being filtered in <strong>the</strong> glomerulus (Chan et al., 1996b), PQ is secreted in<br />

<strong>the</strong> proximal tubule, which is followed by its intracellular accumulation in proximal<br />

tubule cells through an active basolateral uptake mechanism (Chan et al., 1997). Renal<br />

failure proceeds gradually and may produce an unusually rapid rise in serum creatinine<br />

relatively to <strong>the</strong> rise in blood urea nitrogen (low BUN/creatinine ratio) (Chen et al.,<br />

1994a). The observation <strong>of</strong> an unusually high creatinine value in a case <strong>of</strong> upper GIT<br />

bleeding (where one might expect to observe an unusually large increase in BUN but<br />

not creatinine) led to <strong>the</strong> diagnosis <strong>of</strong> PQ toxicity even though <strong>the</strong> patient denied<br />

ingestion. Liver toxicity, as revealed by elevated liver enzymes, jaundice, and<br />

histopathological changes in <strong>the</strong> liver at examination post-mortem, is sometimes seen in<br />

cases <strong>of</strong> poisoning with PQ in humans. The liver lesion caused by PQ displays a picture<br />

<strong>of</strong> centrilobular hepatocellular necrosis and cholestasis and is most <strong>of</strong>ten moderate<br />

(Vale et al., 1987).<br />

6.1.1.2.3 Third phase<br />

Delayed development <strong>of</strong> pulmonary fibrosis is responsible for <strong>the</strong> generally poor<br />

prognosis in acute PQ poisoning. Clinically and radiographically, this appears several<br />

days after ingestion. In <strong>the</strong> typical form, <strong>the</strong> interstitial lesion extends inexorably. The<br />

diagnosis <strong>of</strong> pulmonary fibrosis can, in fact, be made by pulmonary function tests<br />

(PFTs) well before arterial O2 tension (PaO2) decreases (which is a signal <strong>of</strong> a rapid,<br />

fatal clinical evolution). Early gas diffusion disturbances are responsible for alterations<br />

in gases concentrations, which precede radiological manifestations. Radiological lung<br />

changes do not always parallel <strong>the</strong> severity <strong>of</strong> clinical symptoms; <strong>the</strong>y have been<br />

reported to be diffuse, coarse, reticulonodular infiltrates (Bier and Osborne, 1978).<br />

Chest X-ray may be normal, particularly in those patients who die soon after ingestion,<br />

due to multi<strong>organ</strong> failure. More <strong>of</strong>ten, patchy infiltration develops, which may progress<br />

to ground-glass opacification (GGO) <strong>of</strong> one or both lung fields (Vale et al., 1987). The<br />

fibrosis can also be demonstrated on lung scans, where <strong>multiple</strong> reticulated areas adjoin<br />

cystic and tubular lucencies (Im et al., 1991). Im et al. (Im et al., 1991) analyzed<br />

53


Part I - General Introduction __________________________________________________<br />

retrospectively 42 patients with a history <strong>of</strong> PQ ingestion and abnormal findings on<br />

chest radiographs. Radiographic changes during <strong>the</strong> first week after ingestion included<br />

diffuse consolidation (26/39), pneumomediastinum with or without pneumothorax<br />

(15/39), and cardiomegaly with widening <strong>of</strong> <strong>the</strong> superior mediastinum (8/39). Small<br />

cystic and linear shadows began to appear at <strong>the</strong> end <strong>of</strong> <strong>the</strong> first week and represented<br />

<strong>the</strong> preponderant parenchymal abnormality observed after 2-4 weeks. Focal<br />

honeycombing was <strong>the</strong> major parenchymal abnormality after 4 weeks. High-resolution<br />

computed tomography (HRCT) <strong>of</strong> <strong>the</strong> lung 9 months after PQ exposure revealed<br />

localized fibrosis containing small cysts. Pulmonary fibrosis leads to a rapid<br />

development <strong>of</strong> refractory hypoxemia, resulting in death over a period <strong>of</strong> 5 days to<br />

several weeks. Nei<strong>the</strong>r spontaneous nor assisted artificial ventilation can delay <strong>the</strong> fatal<br />

outcome. In <strong>the</strong> final stage, if sepsis (which frequently occurs in <strong>the</strong>se patients, even<br />

when <strong>the</strong>y are treated with antibiotics) does not intervene, <strong>the</strong> PQ poisoning evolves<br />

toward decerebration during mechanical ventilation, with an inspired O2 fraction (FiO2)<br />

<strong>of</strong> 100% and a PaO2 under 30 mmHg. Lee et al. (Lee et al., 1995) reviewed <strong>the</strong> findings<br />

<strong>of</strong> HRCT scans <strong>of</strong> <strong>the</strong> lungs in 16 patients with PQ poisoning. The most common<br />

pattern on initial scans was GGO, present alone or as part <strong>of</strong> a mixed pattern in 13<br />

patients. It was bilateral and diffuse in distribution. Consolidation was present in six<br />

patients, irregular lines in three, and nodules in two patients. On follow-up scans, <strong>the</strong><br />

GGO had changed to consolidation with bronchiectasis. Additional irregular lines and<br />

traction bronchiectasis also were observed. More recently, <strong>the</strong> specific radiologic and<br />

functional sequential changes <strong>of</strong> PQ-<strong>induced</strong> pulmonary damage were well<br />

characterized using HRCT and PFTs in long-term follow-up <strong>of</strong> PQ-poisoned survivals<br />

(Huh et al., 2006). Among <strong>the</strong> cohort <strong>of</strong> 27 patients who had ingested PQ, <strong>the</strong> HRCT<br />

findings showed a normal (n=14) and an abnormal group (n=13). Increased PQ<br />

ingestion in <strong>the</strong> abnormal group was associated with more rapid and severe pulmonary<br />

changes. All <strong>the</strong> patients with normal HRCT findings survived. When <strong>the</strong> serial changes<br />

<strong>of</strong> HRCT are observed, initial GGO indicates primary lung damage from PQ, because<br />

this pattern reflects alveolar edema and inflammatory cell infiltration. GGO on HRCT<br />

peaked on day 7 after ingestion. Between 2 weeks and 1 month, consolidation increased<br />

and pulmonary fibrosis progressed, and slow improvements were observed for up to six<br />

months. Compared with <strong>the</strong> PFTs results obtained at 1 and 6.5 months, expiratory<br />

volume in 1 s (FEV1), forced vital capacity (FVC), and lung carbon monoxide diffusing<br />

capacity (DLCO), all improved slightly. Lung changes after PQ <strong>into</strong>xication may be<br />

54


__________________________________________________Part I - General Introduction<br />

functionally and radiologically reversible following treatment. Although most patients<br />

who have radiological lung changes go on to develop progressive and ultimately fatal<br />

lung damage, <strong>the</strong>re are a few case reports in which patients have developed persistent<br />

radiological changes but have survived (Hudson et al., 1991). There is also evidence<br />

that, in such patients, some recovery may occur over time (Lin et al., 1995; Papiris et<br />

al., 1995).<br />

6.1.1.3 Mild Toxicity<br />

Ingestion <strong>of</strong> PQ ion <strong>of</strong> less than 20-30 mg/Kg produces no symptoms or only mild<br />

GIT symptoms (nausea, irritation and diarrhea) (Vale et al., 1987; Pond, 1990). Renal<br />

and hepatic lesions are ei<strong>the</strong>r minimal or absent. An initial decrease in <strong>the</strong> DLCO is<br />

frequently noted, but development <strong>of</strong> clinical or radiological pulmonary fibrosis is rare.<br />

Full recovery is expected in all cases without sequelae (Vale et al., 1987).<br />

6.1.2 Exposure by dermal route<br />

Although deliberate ingestion or injection is responsible for most cases <strong>of</strong> serious<br />

PQ toxicity, morbidity and mortality can result from o<strong>the</strong>r routes <strong>of</strong> exposure. Indeed,<br />

<strong>the</strong> most important accidental exposure routes for people applying PQ are dermic and<br />

inhalation; in normal use, ingestion is unlikely. Local toxicity is produced by direct<br />

injury to tissues with which <strong>the</strong> herbicide comes <strong>into</strong> contact due to PQ corrosive<br />

effects. Described local effects include skin damage (blistering), as well as nails, nose<br />

and lips ulcers (Samman and Johnston, 1969; Hearn and Keir, 1971; Vale et al., 1987;<br />

Smith, 1988a; H<strong>of</strong>fer and Taitelman, 1989). Contact to concentrated PQ solutions may<br />

cause localized discoloration or a transverse band <strong>of</strong> white discoloration affecting <strong>the</strong><br />

nail plate, although <strong>the</strong> latter may not occur until several weeks after exposure.<br />

Transverse ridging and furrowing <strong>of</strong> <strong>the</strong> nail, progressing to gross irregular deformity <strong>of</strong><br />

<strong>the</strong> nail plate or total loss <strong>of</strong> <strong>the</strong> nail, may also occur (Samman and Johnston, 1969).<br />

Normal nail growth follows. The extent and severity <strong>of</strong> such damage is mainly<br />

dependent on <strong>the</strong> concentration <strong>of</strong> PQ in <strong>the</strong> formulation ra<strong>the</strong>r than <strong>the</strong> dose (as for<br />

GIT lesions). In general, systemic toxicity in humans, after percutaneous exposure,<br />

55


Part I - General Introduction __________________________________________________<br />

seems unusual as reported H<strong>of</strong>fer and Taitelman, 1989b (H<strong>of</strong>fer and Taitelman, 1989)<br />

whom described 15 consecutive cases <strong>of</strong> single exposures <strong>of</strong> <strong>the</strong> skin or eyes during<br />

contact with PQ at working places. From <strong>the</strong>se data it is apparent that a single exposure<br />

<strong>of</strong> healthy skin to PQ solutions only causes local lesions. However, patients with dermal<br />

PQ repeted exposures may have significant skin irritation or can even die. Most <strong>of</strong> <strong>the</strong><br />

fatal cases occurred in developing countries. In all <strong>of</strong> <strong>the</strong>se cases, one or more <strong>of</strong> <strong>the</strong><br />

following factors were present: previous skin damage, caused ei<strong>the</strong>r by PQ itself or by<br />

mechanical or o<strong>the</strong>r chemical means, and prolonged skin contact to clo<strong>the</strong>s soaked in<br />

concentrated PQ, or less concentrated solutions if <strong>the</strong> skin is not washed immediately<br />

after exposure (according to <strong>the</strong> manufacturer’s instructions, correctly diluted spray<br />

solutions should contain no more than 0.05 to 0.2% <strong>of</strong> PQ ion) (Wohlfahrt, 1982). The<br />

lowest known concentration <strong>of</strong> PQ leading to fatal poisoning by dermal route is 5 g/L<br />

(Smith, 1988a). Athanaselis et al. (Athanaselis et al., 1983) reported <strong>the</strong> poisoning <strong>of</strong> a<br />

64-year-old spray operator via <strong>the</strong> skin. Fluid had leaked down his back for several<br />

hours, causing irritation <strong>of</strong> <strong>the</strong> skin. Two days later <strong>the</strong> sprayman visited a doctor, who<br />

advised hospitalization. The patient rejected this advice but was admitted 3 days later<br />

<strong>into</strong> hospital. He died, 12 hours after admission, due to toxic shock, and renal and<br />

respiratory insufficiency. At autopsy, <strong>the</strong> findings were typical <strong>of</strong> PQ poisoning with<br />

fibrosing interstitial pneumonitis and intra-alveolar hemorrhage, renal tubular cell<br />

degeneration, cholestasis, and necrosis <strong>of</strong> <strong>the</strong> back skin. Ano<strong>the</strong>r peculiar case <strong>of</strong> a<br />

fatality from transdermal exposure to PQ was reported in Papua New Guinea (Binns,<br />

1976). The patient, evidently thinking that PQ (20% PQ w/v) would kill lice, applied <strong>the</strong><br />

formulation to his scalp and beard. This produced painful sores and he steadily<br />

deteriorated until dying 6 days after applying <strong>the</strong> PQ to his skin. At autopsy, <strong>the</strong>re were<br />

skin lesions as well as solid and haemorrhagic lungs. Garnier et al. (Garnier et al., 1994)<br />

reported two cases <strong>of</strong> percutaneous exposure. In <strong>the</strong> first case a 36-year-old man applied<br />

20% concentrate to his whole body to cure scabies. He developed extensive ery<strong>the</strong>ma<br />

followed by blistering and 2 days later he was admitted to hospital. He developed<br />

transient renal failure. Dyspnea appeared one week after admission and he deteriorated,<br />

dying 26 days after exposure. In <strong>the</strong> second case, death followed PQ application to<br />

beard and scalp to treat lice (Garnier et al., 1994). An agricultural worker developed<br />

persistent hepatic cholestasis after an episode <strong>of</strong> acute PQ poisoning through skin<br />

absorption (Bataller et al., 2000). Several o<strong>the</strong>r cases <strong>of</strong> percutaneous PQ <strong>into</strong>xication<br />

with respiratory lesions were also reported (Newhouse et al., 1978; Okonek et al., 1983;<br />

56


__________________________________________________Part I - General Introduction<br />

Papiris et al., 1995). A cross-sectional study was undertaken by Castro-Gutierrez et al.<br />

(Castro-Gutierrez et al., 1997) in Nicaragua in order to evaluate any relationship<br />

between respiratory health and PQ exposure. There was a consistent relationship<br />

between a history <strong>of</strong> skin rashes or burns and <strong>the</strong> prevalence <strong>of</strong> dyspnea.<br />

6.1.3 Ocular irritation<br />

Direct eye contact with concentrated solutions will produce caustic ocular injury<br />

dependent on contact time and concentration. Ocular exposure may produce severe<br />

corneal and conjunctival injury and anterior uveitis (Cant and Lewis, 1968). Local<br />

effects to <strong>the</strong> eye may heal only slowly and with scarring (Nirei et al., 1993; McKeag et<br />

al., 2002). McKeag et al. (McKeag et al., 2002) described <strong>the</strong> clinical appearance and<br />

progress <strong>of</strong> bilateral ocular injury caused by PQ on a 69 year old fruit farmer, who<br />

splashed a 20% solution <strong>of</strong> PQ <strong>into</strong> both his eyes. Cingolani et al. (Cingolani et al.,<br />

2006) investigated <strong>the</strong> effects <strong>of</strong> PQ in mice retina. There was no significant decline in<br />

electroretinogram (ERG) a- or b-wave amplitudes after i.v. injection <strong>of</strong> 1 μL <strong>of</strong> 0.5 mM<br />

PQ in C57BL/6 mice, but loss <strong>of</strong> ERG function occurred after injection <strong>of</strong> <strong>the</strong> same<br />

volume <strong>of</strong> 0.75 or 1 mM PQ. Histology in PQ-injected eyes showed condensation <strong>of</strong><br />

chromatin and thinning <strong>of</strong> <strong>the</strong> inner and outer nuclear layers indicating cell death, and<br />

terminal deoxynucleotidyl transferase-mediated dUTP-biotinide end labeling (TUNEL)<br />

demonstrated that one mechanism <strong>of</strong> cell death was apoptosis. Fluorescence in <strong>the</strong><br />

retina and retinal pigmented epi<strong>the</strong>lium after intraocular injection <strong>of</strong> PQ followed by<br />

perfusion with hydroethidine indicated high levels <strong>of</strong> O2 .- , and oxidative damage was<br />

demonstrated by staining for acrolein and enzyme-linked immunosorbent assay<br />

(ELISA) for carbonyl protein adducts. PQ-<strong>induced</strong> damage to <strong>the</strong> outer nuclear layer<br />

was greater in BALB/c mice than in C57BL/6 mice, suggesting strain differences in <strong>the</strong><br />

oxidative defense system <strong>of</strong> photoreceptors and/or o<strong>the</strong>r modifier genes.<br />

57


Part I - General Introduction __________________________________________________<br />

6.1.4 Exposure by inhalation<br />

The large size <strong>of</strong> PQ droplets produced by most commercial agricultural<br />

spraying equipment (greater than 5 μm) generally precludes serious poisoning by <strong>the</strong><br />

inhalational route (Howard et al., 1981; Wojeck et al., 1983; Senanayake et al., 1993).<br />

6.1.5 Muscle toxicity<br />

Myopathy associated with PQ poisoning was reported for <strong>the</strong> first time by<br />

Saunders et al. in 1985 (Saunders et al., 1985). The examination <strong>of</strong> skeletal muscles<br />

obtained at both <strong>the</strong> biopsy and autopsy, revealed findings <strong>of</strong> extensive degeneration<br />

and fibrosis. Koppel et al. (Koppel et al., 1994) reported, in 1994, that extensive<br />

myonecrosis was observed in a specimen <strong>of</strong> post-mortem intercostal muscle <strong>of</strong> a 52year-old<br />

woman who had ingested an unknown dose <strong>of</strong> PQ and died on <strong>the</strong> 11 th day<br />

after ingestion. Vyver at al. (Van de Vyver et al., 1985) reported a case <strong>of</strong> a patient that<br />

died 5 days after ingestion <strong>of</strong> PQ, where levels were higher in <strong>the</strong> skeletal muscle and<br />

an increase <strong>of</strong> creatinine kinase levels appeared on <strong>the</strong> fourth day after admission. More<br />

recently (Tabata et al., 1999), degeneration <strong>of</strong> skeletal muscle, mainly <strong>of</strong> <strong>the</strong> rectus<br />

abdominis m., psoas major m. and diaphragm were also reported. Laboratory data<br />

revealed that <strong>the</strong> plasma CK values (1796 mU/ml) were highest on <strong>the</strong> 5 th day, after<br />

which <strong>the</strong> levels decreased steadily; however, <strong>the</strong>y were maintained at about 900 mU/ml<br />

even on <strong>the</strong> 8 th day.<br />

6.2 Intoxications during pregnancy<br />

In a fatal case <strong>of</strong> PQ poisoning in a pregnant woman, who developed <strong>the</strong> typical<br />

symptoms and signs <strong>of</strong> PQ poisoning and, at post-mortem, had <strong>the</strong> typical lung<br />

pathology <strong>of</strong> PQ poisoning, <strong>the</strong> fetal lungs were normal (Fennelly et al., 1968).<br />

However, Talbot and Fu (Talbot et al., 1988), who reported <strong>the</strong> clinical cases <strong>of</strong> nine<br />

pregnant women who ingested PQ, measured its levels in maternal, fetal and cord blood<br />

in one case and showed that PQ crosses <strong>the</strong> placenta and is concentrated to levels 4-6<br />

times greater than <strong>the</strong> maternal blood. Amnioscopy in ano<strong>the</strong>r case showed PQ levels in<br />

58


__________________________________________________Part I - General Introduction<br />

amniotic fluid nearly twice that <strong>of</strong> maternal blood. The fetus appears to tolerate<br />

maternal PQ poisoning while it is dependent on <strong>the</strong> maternal circulation. The condition<br />

<strong>of</strong> <strong>the</strong> fetus worsened (developed signs <strong>of</strong> PQ poisoning) at delivery (due to exposure to<br />

atmospheric O2), or in utero if <strong>the</strong> gestational age was greater than 30 weeks. Poor late<br />

gestational survival may be due to <strong>the</strong> fact that type II pneumocytes appear between 28<br />

and 32 weeks <strong>of</strong> gestation. All fetuses died, whe<strong>the</strong>r or not emergency cesarean<br />

operation was carried out. Jeng et al. (Jenq et al., 2005) presented a case <strong>of</strong> moderate<br />

PQ poisoning from suicidal ingestion in a woman in <strong>the</strong> third trimester <strong>of</strong> pregnancy.<br />

Despite initial deterioration <strong>of</strong> renal and liver function, she had a normal spontaneous<br />

delivery <strong>of</strong> a healthy baby girl 14 weeks after <strong>the</strong> exposure. The child reached<br />

developmental milestones normally and appeared healthy and well nourished at age 5.<br />

6.3 Incidents <strong>of</strong> pet animals poisoning<br />

PQ poisoning in pet animals is rare. Never<strong>the</strong>less, from time to time PQ is<br />

reported as <strong>the</strong> causative agent in animal poisoning. Longstaffe et al. (Longstaffe et al.,<br />

1981), for example, reported criminal and accidental poisonings <strong>of</strong> cats and dogs, and<br />

Aleksic-Kovacevic et al. (Aleksic-Kovacevic et al., 2003) reported <strong>the</strong> accidental<br />

poisoning by PQ <strong>of</strong> five German shepherd dogs.<br />

7. PREDICTING HUMAN OUTCOME IN PARAQUAT POISONING<br />

Patients who have strong dermal PQ exposure and all who have ingested PQ<br />

require hospitalization and aggressive <strong>the</strong>rapy. PQ poisoning is one <strong>of</strong> those<br />

<strong>into</strong>xications for which it is possible to predict <strong>the</strong> severity and prognosis for individual<br />

patients using specific laboratory tests and information from <strong>the</strong> medical history.<br />

Successful prediction <strong>of</strong> those who may survive PQ poisoning can prevent<br />

inappropriately aggressive treatments, which are normally elaborated, expensive and<br />

have not clearly improved <strong>the</strong> survival rate (Bismuth et al., 1987; Hampson and Pond,<br />

1988), in those who have no hope <strong>of</strong> survival and those only minimally poisoned.<br />

Possible prognostic factors in PQ poisonings are <strong>the</strong> formulation involved, whe<strong>the</strong>r or<br />

59


Part I - General Introduction __________________________________________________<br />

not it was diluted, <strong>the</strong> amount ingested, <strong>the</strong> time since ingestion, <strong>the</strong> presence or absence<br />

<strong>of</strong> food in <strong>the</strong> gut (time since <strong>the</strong> last meal before PQ ingestion), whe<strong>the</strong>r spontaneous<br />

emesis has occurred (<strong>the</strong> colour <strong>of</strong> <strong>the</strong> vomits), treatment already administered,<br />

particularly decontamination measures, and plasma and urinary PQ concentrations<br />

(Proudfoot et al., 1979; Bismuth et al., 1982; Hart et al., 1984; Scherrmann et al., 1987;<br />

Ikebuchi et al., 1993; Proudfoot, 1995). Suicidal PQ poisonings are generally more<br />

severe than accidental poisonings due to higher ingestions. A recent meal, which delays<br />

and reduces absorption, can improve prognosis (Bismuth et al., 1982). A person with<br />

acute PQ ingestion is likely to come to <strong>the</strong> emergency department initially complaining<br />

only <strong>of</strong> an acute corrosive injury, so <strong>the</strong> differential diagnosis should encompass all<br />

corrosive agents. A careful physical examination should include search <strong>of</strong> oral, skin, or<br />

mucous membrane lesions. The endoscopy visualization <strong>of</strong> significant ulcerations in <strong>the</strong><br />

esophagus or stomach within <strong>the</strong> first 24 hours <strong>of</strong> exposure indicates a poor prognosis<br />

(Bismuth et al., 1982). The extent and depth <strong>of</strong> ulceration reflect <strong>the</strong> concentration and<br />

dose <strong>of</strong> PQ that has had contact with <strong>the</strong> mucosal surfaces and, indirectly, <strong>the</strong> amount <strong>of</strong><br />

PQ absorbed systemically. The development <strong>of</strong> renal failure is indicative <strong>of</strong> more severe<br />

toxicity and a worse prognosis than would be predicted for a patient in whom renal<br />

function is preserved (Bismuth et al., 1982; Vale et al., 1987; Baselt and Cravey, 1989).<br />

Almost all patients with renal failure from PQ have significant lung toxicity, but <strong>the</strong>re<br />

are occasional reports <strong>of</strong> renal failure without significant lung toxicity (Dolan et al.,<br />

1984).<br />

The measurement <strong>of</strong> plasma PQ concentration is <strong>the</strong> most reliable method for<br />

assessing <strong>the</strong> prognosis, since <strong>the</strong> severity and rate at which <strong>the</strong> toxic signs evolve<br />

depends on <strong>the</strong> amount <strong>of</strong> PQ absorbed systemically. Intoxicated patients should be<br />

watched and treated expectantly until PQ levels are reported to be nonexistent. PQ<br />

poisoning has been reported in children (McDonagh and Martin, 1970). The clinical<br />

approach for <strong>into</strong>xicated children does not differ from that <strong>of</strong> adults.<br />

7.1 Paraquat quantification in biological samples<br />

Positive semi-quantitative urine tests should be followed by quantitative plasma<br />

and urine PQ levels. Several laboratory analytical methods are available for measuring<br />

PQ in biological samples. PQ has to be detected or quantified in a great variety <strong>of</strong><br />

60


__________________________________________________Part I - General Introduction<br />

biological fluids and tissues, and also in various materials suspected to be <strong>the</strong> source <strong>of</strong><br />

PQ ingestion or exposure. In emergency situations, PQ can be measured in plasma,<br />

urine, gastric aspirate, and dialysates. In <strong>the</strong> field <strong>of</strong> forensic toxicology, PQ can be<br />

assayed in several tissues or in whole blood. The analytical methods for PQ<br />

quantification have been reviewed by Haley, Summers and Scherrmann (Haley, 1979;<br />

Summers, 1980; Scherrmann, 1995). In this review only <strong>the</strong> most applied qualitative<br />

and quantitative tests is described.<br />

7.1.1 Qualitative and semi-quantitative test<br />

The dithionite test is based on <strong>the</strong> reduction <strong>of</strong> PQ by freshly prepared 1%<br />

aqueous sodium dithionite in 0.1 N NaOH to form <strong>the</strong> stable blue radical ion (λmax 603<br />

nm) (Tompsett, 1970) as described above. A visual inspection is immediately<br />

perceptible for <strong>the</strong> presence <strong>of</strong> blue free radical, in comparison with negative and<br />

positive controls. Evaluation <strong>of</strong> <strong>the</strong> colour intensity can be related to a semi-quantitative<br />

scale (see below). The test is rapid, specific and requires <strong>the</strong> availability <strong>of</strong> non-oxidized<br />

sodium dithionite. DQ undergoes a similar reduction to form a yellow green cation (λmax<br />

760 nm).<br />

7.1.2 Quantitative test: spectrophotometry<br />

Probably <strong>the</strong> easiest, rapidest, simplest and <strong>the</strong> method with <strong>the</strong> lowest detection<br />

limit for PQ quantification is based on second or fourth-derivative spectrophotometry<br />

(Fell et al., 1981; Jarvie et al., 1981; Fuke et al., 1992; Kuo et al., 2001). In this mode,<br />

<strong>the</strong> normal absorption spectrum is transformed to <strong>the</strong> second or fourth derivative spectra<br />

<strong>of</strong> <strong>the</strong> PQ cation radical. The sodium dithionite colour reaction is used to detect PQ, and<br />

matrix interference is eliminated by <strong>the</strong> use <strong>of</strong> a chemical deproteinization technique<br />

with sulfosalicylic acid in order to give a clear supernatant, compatible with<br />

spectrophotometry. Derivative spectroscopy confers an advantage over classical<br />

spectrophotometric detection by enhancing <strong>the</strong> PQ •+ peak and suppressing <strong>the</strong> broader<br />

absorption bands resulting from non-specific matrix absorption such as diquat,<br />

haemolysis, bilirubin, or lipemia.<br />

61


Part I - General Introduction __________________________________________________<br />

7.1.2.1 Reagents and <strong>the</strong>ir preparation<br />

62<br />

• A 1.38 mg amount <strong>of</strong> PQ dichloride (Sigma, St. Louis, MO, USA and o<strong>the</strong>r<br />

manufacturers) is dissolved in 1 mL distilled water (1 mg/mL as PQ 2+ );<br />

• Deproteinization reagent: 50g sulfosalicylic acid are dissolved in 100 mL <strong>of</strong><br />

distilled water;<br />

• Alkaline reagent: 40g NaOH are dissolved in 100 mL <strong>of</strong> distilled water;<br />

• Chromogenic reagent: sodium dithionite (Na2S2O4).<br />

7.4.2.2 Analytical conditions<br />

• Instrument: UV-Vis spectrophotometer with a differential analyzing system;<br />

• Cell: plastic-made semimicro-cell with an optical path length <strong>of</strong> 1.0 cm;<br />

7.1.2.3 Procedures<br />

The procedures for urine and blood plasma specimens are detailed below and are<br />

shown in Figure 12:


__________________________________________________Part I - General Introduction<br />

900μL plasma or urine 100μL sulfosalicylicacid<br />

A sodium dithionite<br />

Spatula<br />

(~5 mg)<br />

Centrifugation at 13,000 g, 5 min<br />

800μL <strong>of</strong> supernatant<br />

+<br />

200μL NaOH 10 N<br />

Mix<br />

Spectrophotometer<br />

0<br />

0<br />

Zero-order spectrum Second-derivative spectrum<br />

0.5<br />

(A/Div.)<br />

2<br />

1<br />

403 nm<br />

396 nm<br />

Amplitude<br />

0.2<br />

(A/Div.)<br />

0.02<br />

(A/Div.)<br />

380 440 500 380 440 500<br />

Wavelength (nm)<br />

Wavelength (nm)<br />

A B<br />

Fig. 12 – Pretreatment procedures for <strong>paraquat</strong> in urine and plasma before <strong>the</strong> second<br />

derivative spectrophotometric analysis (A). Zero-order and second-derivative spectrum.<br />

The qualitative analysis is made by observing <strong>the</strong> presence <strong>of</strong> inflection points at about<br />

396 and 403 nm. The quantification is made with amplitudes measurable between 396<br />

and 403 nm (B).<br />

1. 900μL <strong>of</strong> blood plasma and urine are mixed well with 100 µL <strong>of</strong> <strong>the</strong><br />

deproteinization reagent solution;<br />

2. The mixture is centrifuged at 13,000 g for 5 min;<br />

3. 800μL <strong>of</strong> <strong>the</strong> supernatant solution is mixed with 200μL <strong>of</strong> alkaline reagent;<br />

4. A spatula (~5 mg) <strong>of</strong> <strong>the</strong> chromogenic reagent is added to give a blue colour<br />

characteristic <strong>of</strong> <strong>the</strong> PQ .+ .<br />

5. The data <strong>of</strong> a zero-order spectrum is obtained by scanning from 500 to 380 nm<br />

(wavelength space Δλ=0.5 nm) and <strong>the</strong>n second-differentiated (derivative<br />

wavelength space Δλ=4 nm). A qualitative and quantitative analysis <strong>of</strong> PQ .+ is<br />

performed at <strong>the</strong> amplitude peaks <strong>of</strong> 396-403 nm <strong>of</strong> <strong>the</strong> second-derivative<br />

spectrum.<br />

63<br />

2<br />

1


Part I - General Introduction __________________________________________________<br />

64<br />

The calibration curves are constructed by spiking various concentrations <strong>of</strong><br />

<strong>paraquat</strong> <strong>into</strong> blank specimens, and processing in <strong>the</strong> same way as above. The<br />

calibration curve in <strong>the</strong> 0.2-8 μg/mL range obeys Beer’s law. Using <strong>the</strong>se experimental<br />

conditions, <strong>the</strong> intra- and inter-day coefficients <strong>of</strong> variation showed values lower than 5<br />

% and <strong>the</strong> detection limit <strong>of</strong> <strong>the</strong> method was 0.10 μg/mL (Fell et al., 1981; Fuke et al.,<br />

1992; Kuo et al., 2001).<br />

7.2 Predicting <strong>the</strong> outcome from plasma <strong>paraquat</strong> concentrations<br />

The prognosis for a patient with PQ ingestion can be fairly determined by<br />

measuring plasma PQ concentration and its relationship to time <strong>of</strong> ingestion. PQ<br />

concentrations data should be obtained before beginning any treatment that could<br />

decrease <strong>the</strong> levels. A nomogram was initially presented by Proudfoot et al. (Proudfoot<br />

et al., 1979; Proudfoot, 1995) after quantification <strong>of</strong> PQ plasma concentrations at<br />

various times post-ingestion in 79 poisoned victims. Those whose concentrations were<br />

below 2.0, 0.6, 0.3, 0.16, and 0.1 mg/L at 4, 6, 10, 16, and 24 hours after ingestion,<br />

respectively survived. Subsequently this nomogram was refined by Hart et al. (Hart et<br />

al., 1984) by examining plasma PQ concentrations from a larger group <strong>of</strong> patients<br />

(n=218) (Fig. 13). Hart et al. (Hart et al., 1984) produced a contour graph <strong>of</strong> plasma<br />

PQ-to-time relationships for 10, 20, 30, 50, 70, and 90% probability <strong>of</strong> survival. The<br />

50% probability curve reported by Hart et al. (Hart et al., 1984) correlated well with <strong>the</strong><br />

predictive line separating survival from death developed by Proudfoot et al. (Proudfoot<br />

et al., 1979). Hart et al. (Hart et al., 1984) confirmed <strong>the</strong> difficulty <strong>of</strong> Proudfoot et al.<br />

(Proudfoot et al., 1979) in predicting outcome from plasma concentrations data within<br />

<strong>the</strong> first 3 hours. Schermann et al. (Scherrmann et al., 1987) extended this predictive<br />

curve up to <strong>the</strong> 7 th day after <strong>into</strong>xication, and showed that those patients who presented,<br />

within 8 hours, plasma PQ concentrations <strong>of</strong> 10 mg/L or above, usually died from<br />

cardiogenic shock within 24 hours, while those with lower concentrations (but above<br />

<strong>the</strong> predictive line) died <strong>of</strong> pulmonary fibrosis and respiratory failure latter than 24<br />

hours after ingestion.


__________________________________________________Part I - General Introduction<br />

Plasma <strong>paraquat</strong> concentration (µg/mL)<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

0<br />

30%<br />

50%<br />

70%<br />

90%<br />

20%<br />

10%<br />

Probability <strong>of</strong> survival (%)<br />

0 4 8 12 16 20<br />

Hours after swallowing<br />

24 28<br />

Fig. 13 - Nomogram showing relation between plasma <strong>paraquat</strong> concentrations<br />

(μg/mL), time after ingestion, and probability <strong>of</strong> survival. Adapted from Hart et al.<br />

(Hart et al., 1984).<br />

Bismuth et al. (Bismuth et al., 1982) soon confirmed <strong>the</strong> value <strong>of</strong> <strong>the</strong> line with<br />

100% accurate prediction <strong>of</strong> <strong>the</strong> outcome in 17 patients admitted 25 hours or less after<br />

ingestion. Similarly, Schermann et al. (Scherrmann et al., 1987) accurately predicted<br />

<strong>the</strong> outcome in 45 cases. Suzuki et al. (Suzuki et al., 1991) combined <strong>the</strong> data <strong>of</strong><br />

Proudfoot et al. (Proudfoot et al., 1979), Bismuth et al. (Bismuth et al., 1982),<br />

Scherrmann et al. (Scherrmann et al., 1987), and Sawada et al. (Sawada et al., 1988)<br />

with those from a fur<strong>the</strong>r group <strong>of</strong> 78 patients, and concluded that <strong>the</strong> predictive line<br />

correctly identified 101 <strong>of</strong> <strong>the</strong> 102 deaths and 61 <strong>of</strong> <strong>the</strong> 63 survivors evaluated within 24<br />

hours <strong>of</strong> PQ ingestion. Although <strong>the</strong> nomogram can provide a fairly accurate prognosis,<br />

helping in predicting illness severity and death probability if PQ levels can be obtained<br />

immediately, it is inevitable that any predictive line will fail occasionally. Estimation <strong>of</strong><br />

<strong>the</strong> time interval since ingestion is prone to error, particularly during <strong>the</strong> first few hours<br />

when plasma PQ concentrations decline rapidly and a time error <strong>of</strong> even 0.5 to 1.0 hours<br />

may radically alter <strong>the</strong> relationship <strong>of</strong> a concentration to <strong>the</strong> predictive line. In addition,<br />

65


Part I - General Introduction __________________________________________________<br />

plasma PQ concentrations may not be entirely accurate since <strong>the</strong>y may be assayed by<br />

one <strong>of</strong> several methods and publications seldom make clear whe<strong>the</strong>r <strong>the</strong> concentration<br />

reported is that <strong>of</strong> PQ ion or PQ salt, and <strong>the</strong>re is <strong>the</strong> possibility <strong>of</strong> inter-individual<br />

variation in susceptibility to <strong>the</strong> toxic agent, a matter on which <strong>the</strong>re is little, if any,<br />

knowledge.<br />

66<br />

Sawada et al. (Sawada et al., 1988) reported an objective index for <strong>the</strong> prognosis<br />

<strong>of</strong> PQ poisoning based on a study <strong>of</strong> plasma PQ concentrations in 30 patients, 20 <strong>of</strong><br />

whom died and 10 survived. The severity index <strong>of</strong> PQ poisoning (SIPP) is derived from<br />

<strong>the</strong> time (in hours) until <strong>the</strong> beginning <strong>of</strong> treatment from <strong>the</strong> time <strong>of</strong> PQ ingestion,<br />

multiplied by <strong>the</strong> plasma PQ level (μg/mL) on hospital admission.<br />

SIPP = [Plasma level <strong>of</strong> PQ (µg/mL)] × [Time from ingestion to treatment<br />

(hours)])<br />

When <strong>the</strong> SIPP score is less than 10, patients may survive; a score <strong>of</strong> 50 separates<br />

late deaths due to respiratory failure (10


__________________________________________________Part I - General Introduction<br />

study. A discriminant function (D) score> 0.1 predicts survival and D< 0.1 predicts<br />

death. D was calculated as follows:<br />

D=1.3114 - 0.1617 lnT - 0.5408 ln [ln(C × 1000)] where:<br />

T=time since ingestion (h)<br />

C=plasma PQ concentration (μg/mL)<br />

The probability <strong>of</strong> survival was also estimated by Jones et al. (Jones et al., 1999),<br />

who plotted <strong>the</strong> logarithm <strong>of</strong> <strong>the</strong> plasma PQ concentration versus <strong>the</strong> logarithm <strong>of</strong> <strong>the</strong><br />

time since ingestion. The predicted probability <strong>of</strong> survival for any specified time and<br />

concentration was calculated according to <strong>the</strong> following ratio:<br />

exp (logit)/[1 + exp (logit)], where:<br />

logit=0.58–2.33 × log(plasma PQ)–1.15 × log(h since ingestion).<br />

The authors proposed that this equation may be helpful in predicting who will<br />

survive after ingestion <strong>of</strong> PQ up to at least 200 hours after ingestion, and could be used<br />

as a <strong>research</strong> tool for studies on efficacy <strong>of</strong> PQ poisoning treatments.<br />

More recently, Huang et al. (Huang et al., 2003; Huang et al., 2006) successfully<br />

applied <strong>the</strong> Acute Physiology and Chronic Health Evaluation (APACHE) II system<br />

(Knaus et al., 1985) in predicting <strong>the</strong> in-hospital mortality <strong>of</strong> 64 patients with acute PQ<br />

poisoning over a period <strong>of</strong> 12 years. The study demonstrated that <strong>the</strong> APACHE II score<br />

is positively correlated with plasma PQ concentration and with ingested amount <strong>of</strong> PQ.<br />

Non-survivors (n=46) had a higher APACHE II score (23.3 ± 12.7) than survivors<br />

(n=18) (6.1 ± 4.2). All patients who had an APACHE II score greater than 20 died<br />

before discharge. APACHE II score greater than 13 predicted in-hospital mortality with<br />

67% sensitivity and 94% specificity. The authors concluded that <strong>the</strong> APACHE II score<br />

is a simple, reproducible, and practical tool for evaluating <strong>the</strong> severity <strong>of</strong> acute PQ<br />

poisoning.<br />

67


Part I - General Introduction __________________________________________________<br />

7.3 Predicting outcome from urine <strong>paraquat</strong> concentrations<br />

Although plasma PQ concentrations have a greater predictive value, urine data<br />

may contribute to a more rapid evaluation <strong>of</strong> prognosis (Scherrmann et al., 1987). In<br />

addition, certain <strong>of</strong> <strong>the</strong> available urinary tests can be carried out in nearly all hospitals.<br />

A simple urine semi-quantitative test can confirm <strong>the</strong> presence <strong>of</strong> PQ when <strong>the</strong> urine<br />

concentration is about 1.0 μg/mL or greater, by adding 10 mL <strong>of</strong> urine to 2 mL <strong>of</strong> a<br />

freshly prepared 1% sodium dithionite in 1 N sodium hydroxide (Berry and Grove,<br />

1971; Widdop, 1976; Braithwaite, 1987). This qualitative urine test is very easy to<br />

perform, and <strong>the</strong>re is a good correlation between <strong>the</strong> amount <strong>of</strong> PQ present and <strong>the</strong><br />

intensity <strong>of</strong> <strong>the</strong> formed blue colour (Fig. 14). The darker <strong>the</strong> colour, <strong>the</strong> worse is <strong>the</strong><br />

patient’s prognosis (Scherrmann et al., 1987). The patient’s urine should be tested<br />

serially for 24 hours after ingestion. However, a negative result should be interpreted<br />

cautiously, because early urinary semi-quantitative testing may underestimate <strong>the</strong><br />

amount <strong>of</strong> PQ systemically absorbed.<br />

68<br />

0 0.2 0.5 1.0 2.0 8.0 10.0 20.0 100.0<br />

(-) (±) (+) (++) (+++)<br />

(Colourless) (Pale blue) (Light blue) (Navy blue) (Dark blue)<br />

Fig. 14 – Representative qualitative urinary test for <strong>paraquat</strong>. Correlation between <strong>the</strong><br />

PQ concentration (μg/mL) and <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> blue colour change.<br />

Scherrmann et al. (Scherrmann et al., 1987) measured <strong>the</strong> urine PQ concentration<br />

in 53 patients, comparing results with <strong>the</strong> semi-quantitative urine test. Almost all<br />

patients with urinary PQ concentrations less than 1 μg/mL within 24 hours <strong>of</strong> ingestion<br />

survived. Patients with semi-quantitative urine test results showing more than + + (navy<br />

blue; > 10 μg/mL) within 24 hours following ingestion have a high probability <strong>of</strong> death<br />

(Fig. 14 and 15). In contrast, patients showing less than ± (pale blue; < 1 μg/mL) may<br />

survive (Scherrmann et al., 1987). Again, <strong>the</strong>se results should be interpreted with


__________________________________________________Part I - General Introduction<br />

caution since PQ-<strong>induced</strong> acute renal failure influences urine PQ excretion and may<br />

lead to false-negative results.<br />

Urine PQ concentration<br />

(μg/mL)<br />

10000<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

x<br />

x<br />

x x<br />

x<br />

x<br />

x x x<br />

x<br />

x<br />

xx x<br />

x<br />

x x<br />

x x<br />

x<br />

x x<br />

x<br />

x<br />

xx x<br />

x<br />

x<br />

n=53<br />

x<br />

x<br />

x<br />

x x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

6 12 18 24<br />

Dithionite test<br />

Deaths < 24h<br />

Deaths (pulmonary fibrosis) > 24h<br />

Survivors (19)<br />

Fig. 15 – Relationship between urine <strong>paraquat</strong> concentrations and survival. Adapted<br />

from Scherrmann et al. (Scherrmann et al., 1987).<br />

7.4 Additional laboratory tests<br />

Despite <strong>the</strong> fact that plasma concentration is <strong>the</strong> most reliable prognosis factor in<br />

PQ poisoning, its measurement is not readily available in all hospitals. For this reason,<br />

intensive care treatment must <strong>of</strong>ten be undertaken without any information concerning<br />

plasma levels. Neverthless, almost all hospital laboratories can perform usual tests, such<br />

as complete blood count, blood biochemistry, and arterial blood gases. These tests are<br />

generally available immediately after patient admission. Lee et al. (Lee et al., 2002)<br />

reviewed 602 PQ-poisoned patients and reported a correlation between acute death from<br />

PQ poisoning and usual admission laboratory data. These authors concluded that<br />

69


Part I - General Introduction __________________________________________________<br />

besides dermal or inhalational route, and exposure to low PQ quantities, o<strong>the</strong>r factors<br />

are also good prognosis factors, namely young age, lower degrees <strong>of</strong> leukocytosis and<br />

acidosis (<strong>the</strong> non survivors presented lower levels <strong>of</strong> HCO3 - in arterial blood and thus<br />

lower pH), and <strong>the</strong> absence <strong>of</strong> renal, hepatic, and pancreatic failures on admission after<br />

acute PQ poisoning in <strong>multiple</strong> logistic regression analysis. Heart rate, respiratory rate,<br />

hemoglobin, BUN, serum creatinine, aspartate aminotransferase, alanine<br />

aminotransferase, total bilirubin, amylase, and glucose were also significantly lower in<br />

survivors than in nonsurvivors. Therefore, <strong>the</strong> authors concluded that lower differences<br />

between <strong>the</strong> two groups over time may indicate a lower degree <strong>of</strong> PQ exposure or<br />

absorption, or a lower vulnerability to PQ. This supports <strong>the</strong> hypo<strong>the</strong>sis that <strong>the</strong><br />

prognosis in humans is influenced by individual sensitivity (Hart et al., 1984) and<br />

inaccuracies in assessment <strong>of</strong> <strong>the</strong> ingestion-to-presentation interval. Increased levels <strong>of</strong><br />

serum aminotransferases, bilirubin, or amylase were also detected by Bismuth et al.<br />

(Bismuth et al., 1982) in severely poisoned patients. Analysis <strong>of</strong> admission laboratory<br />

data indicated that <strong>the</strong> prognosis <strong>of</strong> patients with acute PQ poisoning depends on renal<br />

function and acid-base balance (Bismuth et al., 1982). White blood cell (WBC) count<br />

at admission is also emphasized as an index <strong>of</strong> predicting outcomes in PQ poisoning by<br />

Kaojarern and Ongphiphadhanakul (Kaojarern and Ongphiphadhanakul, 1991). These<br />

results suggest that certain admission laboratory data may provide as much information<br />

for predicting <strong>the</strong> prognosis as do plasma PQ concentrations.<br />

70<br />

A respiratory index (RI) has been devised to measure pulmonary function trends<br />

in PQ exposures (Suzuki et al., 1989). This may be <strong>of</strong> more value in patients who<br />

present more than 36 hours after PQ ingestion. In a series <strong>of</strong> 51 patients, all 43 patients<br />

with an RI greater than or equal to 1.5 died; all 8 with an RI less than 1.5 survived. This<br />

RI was calculated according to <strong>the</strong> following ratio:<br />

RI=A−aDO2/PO2 , where A−aDO2 is calculated:<br />

713 × FiO2 – PaCO2[FiO2 + (1 - FiO2)/R] – PaO2<br />

The respiratory quotient (R) was assumed to be 0.8 (Bismuth and Hall, 1995).<br />

However, <strong>the</strong> RI is subject to some limitations, and is probably <strong>of</strong> less value than


__________________________________________________Part I - General Introduction<br />

plasma PQ concentrations in early cases than in those who present 36 hours or longer<br />

after ingestion. Fur<strong>the</strong>r assessment <strong>of</strong> <strong>the</strong> RI is required.<br />

Kao et al. (Kao et al., 1999) investigated changes in lung ventilation (LV) and<br />

alveolar permeability (AP) in patients with PQ <strong>into</strong>xication, using<br />

99m Tc<br />

diethylenetriamine pentaacetate (DTPA) radioaerosol lung scintigraphy. Traditional<br />

99m Tc macroaggregated albumin (MAA) perfusion lung imaging was also performed for<br />

comparison. Those patients (69%) with abnormal AP, died. The authors concluded that<br />

AP may help predict outcome in patients with PQ <strong>into</strong>xication.<br />

Analysing <strong>the</strong> serum <strong>of</strong> 21 PQ-poisoned patients, Nakamura et al. (Nakamura et<br />

al., 2001) showed that serum concentrations <strong>of</strong> type IV collagen and tissue inhibitor <strong>of</strong><br />

metalloproteinase-1 (TIMP-1) increased with time in non-survivors but did not change<br />

in survivors. The authors concluded that <strong>the</strong>se parameters may be useful indicators <strong>of</strong><br />

severity and/or prognosis for <strong>the</strong> development <strong>of</strong> respiratory failure in patients with PQ<br />

poisoning.<br />

Finally, pneumoproteinemia, a recent concept in <strong>the</strong> assessment <strong>of</strong> lung diseases,<br />

seems to be promising in predicting <strong>the</strong> PQ outcome; in ARDS several types <strong>of</strong> serum<br />

or bronchoalveolar lavage fluid (BALF) biomarkers, such as surfactant protein (SP)-A, -<br />

B, and –D, and Clara cell (CC)16 have been evaluated with success (Doyle et al., 1995;<br />

Doyle et al., 1998; Hermans and Bernard, 1998; Kuroki et al., 1998; Hermans and<br />

Bernard, 1999). Pan et al. (Pan et al., 2002) observed an increase <strong>of</strong> serum SP-D in rats<br />

exposed (i.p.) to PQ plus O2. These authors proposed that serum SP-D may be a useful<br />

biomarker <strong>of</strong> lung injury and type II cell hyperplasia, at least in rodents, as a<br />

consequence <strong>of</strong> PQ exposure. This innovative approach to evaluate PQ-<strong>induced</strong> lung<br />

injury was only recently investigated in humans by Hantson et al. (Hantson et al.,<br />

2006). These authors described a case report <strong>of</strong> a 20-year-old man, who ingested<br />

100 mL <strong>of</strong> a 20% PQ formulation. Serum CC16 increased gradually with <strong>the</strong><br />

progression <strong>of</strong> renal impairment and serum SP-A and SP-B levels increased before any<br />

significant changes in pulmonary gas exchanges. The SP-A, -B and CC16 levels in<br />

BALF were within normal limits. The immunostaining studies using antibodies (Ab)<br />

directed against CC16, SP-A and -B were performed on post-mortem lung tissue<br />

specimens and showed that <strong>the</strong> labelling for SP-A and -B was reduced or absent<br />

following PQ toxicity, while Clara cells were relatively preserved.<br />

71


Part I - General Introduction __________________________________________________<br />

72<br />

8. TREATMENT<br />

The high toxicity <strong>of</strong> PQ, coupled with its widespread use and ready accessibility,<br />

results in many human exposures, by both unintentional and deliberate self-poisonings.<br />

Unless <strong>the</strong> exposure is negligible, all PQ poisonings require immediate treatment and<br />

close monitoring in a hospital setting. The “window <strong>of</strong> opportunity” for any effective<br />

treatment <strong>of</strong> <strong>paraquat</strong> poisoning is very short, only a few hours at most. All attempts<br />

should be made to obtain an accurate history for any agrochemical exposure.<br />

In view <strong>of</strong> <strong>the</strong> proposed <strong>mechanisms</strong> <strong>of</strong> PQ toxicity, it has been possible, at<br />

several points, to interrupt <strong>the</strong> toxic pathway. Management has been directed primarily<br />

at removing PQ from <strong>the</strong> GIT (preventing absorption), increasing its excretion from<br />

blood, and, traditionally, by taking measures aimed at preventing pulmonary damage<br />

with anti-inflammatory agents and some newer drugs. In Figure 16, a flowchart guide<br />

currently used in <strong>the</strong> management <strong>of</strong> poisoned patients is presented.<br />

8.1 Preventing <strong>paraquat</strong> absorption<br />

The key to successful treatment <strong>of</strong> an acute PQ exposure depends almost entirely<br />

on aggressive early decontamination measures to limit absorption. If <strong>the</strong>re has been<br />

dermal exposure, ei<strong>the</strong>r primarily or secondarily from contact with contaminated<br />

vomitus, <strong>the</strong> clothing should be removed immediately and <strong>the</strong> skin washed gently but<br />

thoroughly with soap and water to prevent transdermal absorption. Harsh scrubbing<br />

should not be conducted because <strong>the</strong> resultant skin abrasion could actually increase <strong>the</strong><br />

transdermal absorption <strong>of</strong> PQ. PQ-exposed eyes should be irrigated with copious<br />

amounts <strong>of</strong> tepid water or normal saline for at least 15 to 20 min. Since PQ avidly binds<br />

to clay, oral administration <strong>of</strong> mineral adsorbents may be useful as a pre-hospital<br />

treatment for minimizing PQ absorption. Measures to limit absorption that have been<br />

employed include <strong>the</strong> addition <strong>of</strong> an emetic to all PQ formulations, induction <strong>of</strong> emesis<br />

with syrup <strong>of</strong> ipecac, whole gut lavage and oral administration <strong>of</strong> mineral adsorbent.


__________________________________________________Part I - General Introduction<br />

Poisoning<br />

confirmation<br />

Preventing<br />

GIT<br />

absorption<br />

Preventing<br />

pulmonary<br />

damage<br />

A. Skin and eye decontamination<br />

• Flush <strong>the</strong> skin immediately with copious amounts <strong>of</strong> water and seek subsequent treatment by a<br />

dermatologist 1 ;<br />

• Eyes – irrigated at least 15 min with clean water, local antibiotics to prevent secondary infection and<br />

seek subsequent treatment by an ophthalmologist.<br />

B. Inhalation<br />

• No specific treatment is necessary o<strong>the</strong>r than symptomatic for epistaxis. No need to perform urine tests.<br />

When applied as recommended <strong>the</strong> droplets are too large to be inhaled <strong>into</strong> <strong>the</strong> lungs;<br />

C. Significant PQ ingestion suspected on<br />

history and/or examination<br />

• Gastric lavage<br />

(< 2 hour after<br />

ingestion 2 )<br />

+<br />

Vomiting?<br />

NO<br />

• Activated charcoal<br />

(>12 yrs: 100 g/0.5 L water 3 ;<br />

12 yrs: 100-150g,<br />

12 yrs: 100-150,<br />

12 yrs: 10-15 mg s.c.<br />

every 4 hours, 3000/m 3 , 1 day;<br />

AGGRESSIVE THERAPY<br />

CHP (4 days), 7 sessions<br />

(6-8 hours each). Calcium and platelets<br />

must be replenished if depleted<br />

DFO 100 mg/Kg in 5% destrose<br />

solution, 500 mL, infused over<br />

24 hours (21 mL/h),<br />

only after <strong>the</strong> first CHP<br />

PQ semi-quantitative<br />

test - urine or gastric aspirate (10 mL sample + 2 mL <strong>of</strong> freshly<br />

prepared 1% sodium dithionite in 1N NaOH)<br />

SEE BOX BELOW<br />

OR<br />

If negative<br />

Vitamin-E<br />

300 mg<br />

per os<br />

twice daily<br />

If positive If positive<br />

Plasma 7 <strong>paraquat</strong> level<br />

(taken between 5 and 24 hours)<br />

+<br />

Repeat at 6<br />

hours<br />

Nomogram for prognosis<br />

•Sorbitol<br />

solution 70%<br />

(> 12yrs: 1-2 mL/Kg<br />


Part I - General Introduction __________________________________________________<br />

Fig. 16 - Flowchart guide usually followed for <strong>the</strong> management <strong>of</strong> <strong>paraquat</strong> poisoning.<br />

PQ, <strong>paraquat</strong>; i.v., intravenous; PaO2, partial pressure <strong>of</strong> oxygen in arterial blood; O2,<br />

oxygen; NO, nitric oxide; CHP, charcoal hemoperfusion; CP, cyclophosphamide; MP,<br />

methylprednisolone; DFO, desferoxamine; NAC, N-acetylcysteine; DEX,<br />

dexamethasone; WBC, white-blood-cells; 1 If systemic toxicity is suspected, test urine<br />

for PQ. There is little data for time to peak plasma levels by skin absorption, but if <strong>the</strong><br />

urine is negative for 24 hours, systemic toxicity can probably be disregarded. If <strong>the</strong><br />

urine test is positive or if <strong>the</strong>re is any doubt about potential systemic toxicity, assess<br />

blood concentrations and treat for systemic toxicity as described for ingestion; 2 Risk <strong>of</strong><br />

inducing bleeding, perforation or scarring due to additional trauma to fragilized tissues.<br />

Gastric lavage without administration <strong>of</strong> an adsorbent has not shown any clinical<br />

benefit; 3 Or in 250 mL <strong>of</strong> cathartics (it will increase gut motility to improve excretion <strong>of</strong><br />

<strong>the</strong> charcoal-PQ complex) via nasogastric tube; 4 Maximum dose is 50 g; 5 Repeat doses<br />

<strong>of</strong> cathartics may result in fluid and electrolyte imbalances, particularly in children, and<br />

are <strong>the</strong>refore not recommended; 6 Particularly important as a mean to correct<br />

dehydration, accelerating excretion, reducing tubular concentrations and correcting any<br />

metabolic acidosis. However, fluid balance must be monitored to avoid fluid overload if<br />

renal failure develops. In this case hemodialysis or hem<strong>of</strong>iltration may be required;<br />

7<br />

Plasma should be analyzed ra<strong>the</strong>r than serum, because serum PQ concentrations are<br />

approximately 3 fold lower than those in plasma obtained from <strong>the</strong> same blood sample.<br />

If only serum is available results should be interpreted with caution in relation to<br />

survival curves. Plasma should be stored in plastic and not in glass tubes because PQ 2+<br />

adsorb onto glass surfaces.<br />

In 1975, <strong>the</strong> manufacturer <strong>of</strong> PQ (ICI) added a potent emetic, PP796 (a<br />

phosphodiesterase inhibitor), to liquid and solid PQ formulations. There are a few<br />

published laboratory experiments reporting <strong>the</strong> use <strong>of</strong> emetic-containing PQ<br />

formulations. A study in rats suggested that <strong>the</strong> emetic used in proprietary PQ<br />

preparations may itself possess cardio-respiratory toxicity when given i.v., although <strong>the</strong><br />

relevance <strong>of</strong> this finding to human PQ ingestions is uncertain (Noguchi et al., 1985).<br />

Unfortunately, despite <strong>the</strong> occurrence <strong>of</strong> earlier vomiting, Bramley and Hart (Bramley<br />

and Hart, 1983) were unable to demonstrate an improved prognosis in patients who had<br />

ingested emetic-containing, ra<strong>the</strong>r than non-emetic-containing PQ formulations.<br />

Subsequent reports (Onyon and Volans, 1987) from <strong>the</strong> same study have also failed to<br />

74


__________________________________________________Part I - General Introduction<br />

record any significant reduction in mortality since <strong>the</strong> introduction <strong>of</strong> <strong>the</strong> emetic PP796.<br />

A decrease in <strong>the</strong> mortality <strong>of</strong> PQ poisoning as a result <strong>of</strong> <strong>the</strong> introduction <strong>of</strong> <strong>the</strong> emeticcontaining<br />

formulation also failed to be noted by o<strong>the</strong>r investigators (Bismuth et al.,<br />

1982). No clinical or experimental studies involving <strong>the</strong> use <strong>of</strong> ipecac syrup in PQ<br />

poisoning have been reported. Syrup <strong>of</strong> ipecac may be <strong>of</strong> value in a home setting if<br />

immediately available. If ipecac is used, it should be administered within one hour and<br />

only in an alert conscious patient. The risk <strong>of</strong> worsening <strong>the</strong> GIT caustic injury must be<br />

balanced against <strong>the</strong> lethality <strong>of</strong> <strong>the</strong> amount ingested.<br />

There have been only two clinical studies published where <strong>the</strong> authors made<br />

specific mention to <strong>the</strong> efficacy <strong>of</strong> gastric lavage. Bismuth et al. (Bismuth et al., 1982)<br />

were not able to establish <strong>the</strong> value <strong>of</strong> gastric lavage in a review involving 28 patients.<br />

Bramley and Hart (Bramley and Hart, 1983) were unable to demonstrate an improved<br />

prognosis resulting from <strong>the</strong> use <strong>of</strong> gastric lavage in a study <strong>of</strong> 262 cases <strong>of</strong> PQ<br />

poisoning referred to a poison information service in <strong>the</strong> United Kingdom. Without<br />

administration <strong>of</strong> an adsorbent, gastric lavage should never be used. The use <strong>of</strong><br />

sterilized diatomaceous clays (bentonite and Fuller’s Earth) in <strong>the</strong> gastric lavage<br />

solution has been performed prior to <strong>the</strong>ir continued GIT administration (along with <strong>the</strong><br />

cathartic magnesium sulfateate) (Meredith and Vale, 1987; Vale et al., 1987). There are<br />

additional <strong>the</strong>oretical objections to gastric lavage following PQ ingestion. Ulceration <strong>of</strong><br />

<strong>the</strong> oropharyngeal and esophagogastric mucosal surfaces by concentrated PQ<br />

formulations is likely to make <strong>the</strong> procedure hazardous and risk <strong>of</strong> fur<strong>the</strong>r injury exists<br />

(Meredith and Vale, 1987).<br />

Although Fuller’s Earth and bentonite are recommended as adsorbents<br />

(administered orally or via nasogastric tube) in PQ ingestions, <strong>the</strong> ready availability and<br />

<strong>the</strong> equal if not greater efficacy <strong>of</strong> activated charcoal to bind PQ make it <strong>the</strong> agent <strong>of</strong><br />

choice (Okonek et al., 1976; Okonek et al., 1982a; Okonek et al., 1982b). Activated<br />

charcoal in suspension with magnesium citrate effectively adsorbs PQ, an effect that is<br />

maximal at pH 7.8 as observed in vitro and in vivo by <strong>the</strong> improvement <strong>of</strong> rats survival<br />

rate to 94% in opposition to 63% in <strong>the</strong> activated charcoal and Fuller’s Earth groups,<br />

and 69% in <strong>the</strong> magnesium citrate group (Gaudreault et al., 1985). Activated charcoal,<br />

100 g for adults and 2 g/Kg BW for children, should be given unless <strong>the</strong>re is a<br />

contraindication, such as protracted vomiting or severe burns <strong>of</strong> <strong>the</strong> oral mucous<br />

membranes. Rapid control <strong>of</strong> repeated vomiting with antiemetics and promotility agents<br />

is essential when <strong>the</strong> patient cannot retain <strong>the</strong> adsorbent. A total <strong>of</strong> three doses <strong>of</strong><br />

75


Part I - General Introduction __________________________________________________<br />

activated charcoal at 2-hour intervals have been suggested. The airway should be<br />

protected appropriately to prevent aspiration <strong>of</strong> gastric contents.<br />

76<br />

Yamashita et al. (Yamashita et al., 1987) have reported <strong>the</strong> results <strong>of</strong> gastric and<br />

intestinal lavage with <strong>the</strong> cation-exchange-resin, Kayexalate, in PQ-poisoned patients.<br />

Six <strong>of</strong> 11 patients treated in this manner survived, while 11 patients who did not receive<br />

Kayexalate died. Unfortunately, it is not possible to judge whe<strong>the</strong>r <strong>the</strong> severity <strong>of</strong><br />

poisoning was comparable in <strong>the</strong> two groups <strong>of</strong> patients because blood PQ<br />

concentrations were not provided.<br />

PQ has very low bioavailability but peak concentrations occur very early. Thus,<br />

<strong>the</strong>se procedures to prevent absorption are only likely to work if given very soon after<br />

poisoning (within 1-2 hours). In practice, however, <strong>the</strong>y are used very frequently<br />

irrespective <strong>of</strong> <strong>the</strong> delay between poisoning and treatment.<br />

8.2 Increasing <strong>paraquat</strong> elimination<br />

In cases <strong>of</strong> PQ poisoning by ingestion once GIT decontamination has been<br />

performed, <strong>the</strong>re remain two additional treatment strategies. The first is to attempt to<br />

alter <strong>the</strong> herbicide's toxicokinetics (i.e., its distribution in <strong>the</strong> body after ingestion). The<br />

second is to attempt to modify its toxicodynamics (i.e., <strong>the</strong> herbicide’s effects on <strong>the</strong><br />

target <strong>organ</strong>s).<br />

8.2.1 Extracorporeal elimination<br />

The goal <strong>of</strong> extracorporeal elimination procedures is to remove PQ from <strong>the</strong><br />

circulation and prevent its uptake by pneumocytes and Clara cells. The only method that<br />

has been shown to be efficient and to enhance <strong>the</strong> extracorporeal elimination <strong>of</strong> PQ is<br />

charcoal hemoperfusion (CHP) with CLPQ values that may be as high as 170 mL/min.<br />

These high clearances, however, do not allow extrapolation <strong>of</strong> <strong>the</strong> efficacy <strong>of</strong> <strong>the</strong>se<br />

procedures in removing clinically significant quantities <strong>of</strong> PQ, as <strong>the</strong> amount removed<br />

depends on <strong>the</strong> plasma level, which always decreases rapidly. A 6–8-hours course <strong>of</strong><br />

CHP may be beneficial if <strong>the</strong> procedure can be instituted within 4 hours after ingestion<br />

and its efficacy is maintained even when plasma concentration is less than 0.2 mg/L.


__________________________________________________Part I - General Introduction<br />

Although <strong>the</strong> plasma PQ levels peak early in <strong>the</strong> course <strong>of</strong> <strong>the</strong> poisoning, in view <strong>of</strong> <strong>the</strong><br />

mechanism <strong>of</strong> PQ toxicity, additional efficacy <strong>of</strong> CHP may persist even within 10 to 12<br />

hours after ingestion, before <strong>the</strong> absorbed PQ is extensively distributed to <strong>the</strong> tissues<br />

(Smith, 1987). Most toxicologists currently recommend rapid initiation <strong>of</strong> CHP to lower<br />

plasma PQ levels and to limit pulmonary and o<strong>the</strong>r <strong>organ</strong> uptake <strong>of</strong> PQ. Patients with<br />

plasma levels <strong>of</strong> 3 mg/L or greater should probably not be considered for CHP<br />

treatment because <strong>of</strong> <strong>the</strong> uniformly poor prognosis and a lack <strong>of</strong> demonstrated efficacy<br />

for <strong>the</strong> procedure (Hampson and Pond, 1988). Fur<strong>the</strong>rmore, <strong>the</strong> renal CLPQ with normal<br />

kidneys is 3-10 times more efficient than CLPQ by means <strong>of</strong> CHP (Proudfoot et al.,<br />

1987). Tabei et al. (Tabei et al., 1982) studied, in vitro and in vivo, <strong>the</strong> efficiency <strong>of</strong><br />

CHP for <strong>the</strong> removal <strong>of</strong> PQ. At a flow rate <strong>of</strong> 200 mL/min, 93-99% <strong>of</strong> PQ in 4 L <strong>of</strong><br />

solution (5, 10, 100 ppm) was removed in less than 160 min. The elimination t1/2 was 16<br />

min and 10 sec. At 100 mL/min, it was 49 min 30 sec. Of 23 PQ poisoning cases, 15<br />

patients underwent CHP, <strong>of</strong> which 10 died <strong>of</strong> respiratory failure within 28 days (7.6 ±<br />

2.9) and 5 survived without pulmonary complications. Of eight patients who did not<br />

receive CHP, six died <strong>of</strong> respiratory failure within 97 days (33.4 ± 18.8), even when<br />

<strong>the</strong>ir general condition was good upon admission. In one patient, whose PQ<br />

concentration in blood was followed, 99% <strong>of</strong> <strong>the</strong> PQ was removed from circulating<br />

blood by a single CHP. They concluded that CHP is effective for <strong>the</strong> removal <strong>of</strong> PQ<br />

from blood in vivo and from solution in vitro. CHP may thus improve survival after PQ<br />

ingestion. Analysing 105 patients who had swallowed one to three mouthfuls <strong>of</strong> PQ<br />

solution (24.5% w/v) Hong et al. (Hong et al., 2003) also concluded that adequate CHP<br />

appears to be an indispensable treatment for patients with acute PQ poisoning. When<br />

CHP is effective (i.e, when PQ levels in venous outlet approximates to zero), <strong>the</strong> plasma<br />

PQ concentration drops dramatically within 1-3 hours. Unless <strong>the</strong> procedure is begun at<br />

an early stage, when PQ is concentrated in <strong>the</strong> central compartment, a poor total body<br />

CLPQ by extracorporeal techniques and a rise in plasma concentrations for several hours<br />

following completion <strong>of</strong> CHP, may ensue, which can be explained by extensive PQ<br />

tissue distribution (a rebound in plasma concentrations is observed) and its slow<br />

redistribution back <strong>into</strong> <strong>the</strong> circulation following termination <strong>of</strong> <strong>the</strong> extracorporeal<br />

procedure (De Broe et al., 1986). Because <strong>of</strong> <strong>the</strong>se factors, Okonek et al. (Okonek et al.,<br />

1979; Okonek et al., 1982b) proposed that “continuous” (repeated) CHP should be<br />

performed. However, treating PQ poisoning with CHP has been <strong>the</strong> subject <strong>of</strong><br />

considerable controversy, <strong>the</strong> weight <strong>of</strong> evidence in <strong>the</strong> published literature showing a<br />

77


Part I - General Introduction __________________________________________________<br />

lack <strong>of</strong> clinical efficacy in several cases (Bismuth et al., 1982; Castro et al., 2005). CHP<br />

is also <strong>of</strong>ten complicated by thrombocytopenia, as platelets adhere to <strong>the</strong> cartridge<br />

(Winchester, 2002).<br />

78<br />

Serial and combined CHP with hemodialysis (HD) have also been recommended,<br />

particularly during <strong>the</strong> first 24 hours after exposure (Proudfoot et al., 1979; Tabei et al.,<br />

1982). The CLPQ achieved with HD is good when PQ plasma levels are high (around 10<br />

mg/L). However, CLPQ drops remarkably when <strong>the</strong> plasma concentration is less than 1<br />

mg/L. The CLPQ achievable by hemodialysis (HD) can be as high as 150 mL/min<br />

(Okonek et al., 1982b). However, considering that <strong>the</strong> plasma PQ concentration is<br />

usually relatively low, <strong>the</strong> actual amount <strong>of</strong> PQ removed by this extracorporeal<br />

procedure may be clinically insignificant compared with <strong>the</strong> ingested dose (Proudfoot et<br />

al., 1987). In addition, <strong>the</strong> CLPQ with HD decreases considerably when <strong>the</strong> plasma<br />

concentration falls to less than 0.5 mg/L. Neverthless, HD should be used, when<br />

indicated, for <strong>the</strong> treatment <strong>of</strong> PQ-<strong>induced</strong> renal failure.<br />

Plasmapheresis was also tentatively used in acute PQ poisonings, though with no<br />

success (Tsatsakis et al., 1996).<br />

8.2.2 Forced diuresis and peritoneal dialysis<br />

O<strong>the</strong>r <strong>the</strong>rapies that have been investigated include removal <strong>of</strong> PQ from <strong>the</strong> blood<br />

by forced diuresis and peritoneal dialysis. Forced diuresis was initially popular in <strong>the</strong><br />

management <strong>of</strong> patients with PQ poisoning (Kerr et al., 1968; Fennelly et al., 1971).<br />

However, forced diuresis is not very effective since PQ tubular reabsorption is small<br />

(Lock and Ishmael, 1979). Moreover, pulmonary edema, a complication <strong>of</strong> PQ<br />

poisoning as well as <strong>of</strong> forced diuresis, increases morbidity and makes patient<br />

management more difficult. Fluid replacement must be undertaken with careful<br />

monitoring <strong>of</strong> respiratory function and urine output. Never<strong>the</strong>less, furosemide and <strong>the</strong><br />

early administration <strong>of</strong> i.v. fluids and electrolytes, to maintain adequate urine flow<br />

(achieving an urine output <strong>of</strong> 1 to 2 mL/Kg/hour), are important because <strong>the</strong> kidneys are<br />

<strong>the</strong> major physiological route for PQ excretion. A brisk urine flow supports both<br />

glomerular filtration and tubular secretion <strong>of</strong> PQ and delays <strong>the</strong> onset <strong>of</strong> an acute<br />

oliguric renal failure. PQ may cause peripheral vasodilatation with "third spacing"<br />

(Webb and Leopold, 1983) and intra-renal vasoconstriction (Lock and Ishmael, 1979).


__________________________________________________Part I - General Introduction<br />

These <strong>mechanisms</strong> account for a functional component <strong>of</strong> <strong>the</strong> early stages <strong>of</strong> PQ-<br />

<strong>induced</strong> renal failure, which is mostly reversible. Unfortunately, this may occur in <strong>the</strong><br />

first several hours following PQ poisoning, with decreases <strong>of</strong> CLPQ toge<strong>the</strong>r with<br />

increases <strong>of</strong> its t1/2 and consequent generation <strong>of</strong> lethal concentrations <strong>of</strong> PQ in <strong>the</strong> lung<br />

tissue (Bismuth et al., 1987). Bismuth et al. (Bismuth et al., 1982) suggested that<br />

functional renal failure has no prognostic value, while <strong>organ</strong>ic renal failure (proximal<br />

acute renal tubular necrosis) has greater importance in predicting <strong>the</strong> outcome from PQ<br />

poisoning. This functional renal impairment should be promptly corrected with i.v.<br />

volume expansion to allow maximal renal excretion <strong>of</strong> systemically absorbed PQ before<br />

<strong>the</strong> onset <strong>of</strong> renal tubular necrosis from direct action <strong>of</strong> <strong>the</strong> herbicide on <strong>the</strong> kidney<br />

(Lock and Ishmael, 1979).<br />

Peritoneal dialysis is a poor mean <strong>of</strong> removing PQ (Carson, 1972). Indeed, this<br />

technique is only able to eliminate small quantities <strong>of</strong> <strong>the</strong> herbicide when plasma levels<br />

are very high.<br />

8.3 Supportive <strong>the</strong>rapies<br />

Patients poisoned with PQ are always dehydrated to some extent due to GIT fluid<br />

losses (Webb and Leopold, 1983; Williams et al., 1984). Besides maintaining kidney<br />

perfusion, early i.v. fluids and electrolytes administration are also important to correct<br />

dehydration.<br />

Reduction <strong>of</strong> O2 supply (hypooxygenation) has been tried because <strong>of</strong> evidence in<br />

animals <strong>of</strong> a relationship between <strong>the</strong> FiO2 and <strong>the</strong> severity <strong>of</strong> <strong>the</strong> pulmonary damage<br />

(Fisher et al., 1973; Rhodes et al., 1976; Kehrer et al., 1979). Animals poisoned with<br />

PQ die more quickly in an O2-enriched atmosphere than do poisoned animals breathing<br />

room air (Fisher et al., 1973; Kehrer et al., 1979). Similarly, poisoned animals kept in a<br />

somewhat hypoxic atmosphere had lower mortality rates than animals kept in room air<br />

(Rhodes et al., 1976). Oxygen may increase lung injury by providing additional<br />

substrate for O2 .- formation. In humans, O2 also appears to accelerate lung damage, and<br />

thus artificial ventilation with low O2 concentrations (


Part I - General Introduction __________________________________________________<br />

pressure (PEEP) and continuous positive pressure breathing. Supplemental O2 is given<br />

when necessary for symptomatic relief, but mechanical ventilation would not be <strong>of</strong>fered<br />

as a treatment option for a patient who is obviously in impending respiratory failure due<br />

to pulmonary fibrosis.<br />

80<br />

Relief <strong>of</strong> pain and anxiety is essential. Because medical <strong>the</strong>rapy is so highly<br />

unsuccessful in reversing moderate to severe PQ ingestions, <strong>the</strong> health care providers,<br />

<strong>the</strong>ir patients, and <strong>the</strong> patients’ families are <strong>of</strong>ten bewildered. The art <strong>of</strong> medicine is<br />

crucial here as a multidisciplinary support. Honesty about prognosis, without taking<br />

away hope, and emphasizing what can be done (i.e., pain relief and pastoral and social<br />

service care) are keystone approaches to a grim situation (Vale et al., 1987).<br />

First attempted in 1968 (Mat<strong>the</strong>w et al., 1968), lung transplantation has been used<br />

in highly selected patients but mostly with unsuccessful outcomes (Saunders et al.,<br />

1985). Since muscles are important body reservoirs for PQ, <strong>the</strong> herbicide release from<br />

muscles may occur when weaning from mechanical ventilation is started resulting in a<br />

new lung injury.<br />

8.4 Measures to prevent lung damage<br />

Research <strong>into</strong> <strong>the</strong> <strong>mechanisms</strong> <strong>of</strong> PQ toxicity and <strong>the</strong> development <strong>of</strong> antidotes,<br />

although very productive in terms <strong>of</strong> <strong>the</strong> information gained about free radical-mediated<br />

toxicity and <strong>the</strong> polyamines and <strong>the</strong>ir uptake pathways (Smith, 1988b), has yielded little<br />

hope for PQ-poisoned patients. No antidote has been currently recommended on <strong>the</strong><br />

basis <strong>of</strong> convincing evidence obtained in patients. Certain <strong>of</strong> <strong>the</strong>se compounds appear to<br />

give promising results in experimental animals, partly because <strong>the</strong>y are administered<br />

ei<strong>the</strong>r prophylactically or early in <strong>the</strong> course <strong>of</strong> <strong>the</strong> poisoning. Confirmation <strong>of</strong> efficacy<br />

is not available from patient clinical data because <strong>of</strong> <strong>the</strong> heterogeneous nature <strong>of</strong> <strong>the</strong><br />

patient populations, a lack <strong>of</strong> prospective, controlled trials without numerous<br />

confounding variables, and <strong>the</strong> fact that patients may be presented at emergency rooms<br />

after most <strong>of</strong> <strong>the</strong> PQ has been eliminated from <strong>the</strong> body.<br />

As a first approach, most potential antidotes have been directed toward<br />

compounds that detoxify <strong>the</strong> O2 .- or <strong>the</strong> o<strong>the</strong>r subsequently formed ROS. These have<br />

included:


__________________________________________________Part I - General Introduction<br />

Superoxide dismutase or mimetic enzymes<br />

Under normal circumstances, O2 .- produced by PQ and o<strong>the</strong>r chemicals is kept<br />

under control by <strong>the</strong> superoxide dismutase (SOD) enzymes. The use <strong>of</strong> SOD as a<br />

treatment to ameliorate PQ-<strong>induced</strong> injuries has produced variable results.<br />

Exogenously-administered SOD conferred protection in young rats that had been<br />

challenged with PQ (Autor, 1977). Also, in adult rats, SOD reduced <strong>the</strong> mortality to PQ<br />

challenge from ≈80 to 45% over a 28-day period (Wasserman and Block, 1978). In<br />

contrast, <strong>the</strong> results from most o<strong>the</strong>r studies in which SOD had been employed as an<br />

antioxidant treatment for PQ toxicity demonstrate that when SOD was administered by<br />

continuous i.v. infusion, it failed to ameliorate <strong>the</strong> toxic effects <strong>of</strong> <strong>the</strong> herbicide (Block,<br />

1979). In addition, SOD administered by <strong>the</strong> parenteral route was not effective in<br />

human poisonings (Fairshter et al., 1979). Although <strong>the</strong>se differences are not easily<br />

explained, it has been reported that <strong>the</strong> lack <strong>of</strong> SOD effectiveness in protecting against<br />

PQ toxicity can be attributed to its physicochemical properties; this enzyme cannot<br />

enter <strong>the</strong> target cell membrane because <strong>of</strong> its high molecular size (which prevents<br />

intracellular transport) or its charge (which prevents its adherence to targets) (Freeman<br />

et al., 1985). More recently, in order to circumvent <strong>the</strong>se problems, investigators have<br />

used low-molecular-weight metalloporphyrin SOD mimetics or liposomal encapsulated<br />

SOD for <strong>the</strong> purpose <strong>of</strong> successfully treating oxidative stress-<strong>induced</strong> injuries. More<br />

precisely, Day and Crapo (Day and Crapo, 1996) employed <strong>the</strong> low-molecular-weight<br />

metalloporphyrin SOD mimetic, tetrakis-(4-benzoic acid) porphyrin (MnTBAP), to<br />

protect mice against PQ-<strong>induced</strong> lung injury. This SOD mimetic has been demonstrated<br />

to penetrate cell membranes, retain its intracellular activity and also protect endo<strong>the</strong>lial<br />

cells against intracellular PQ-<strong>induced</strong> injury in vitro (Day and Crapo, 1996). However,<br />

no studies have examined <strong>the</strong> role <strong>of</strong> liposomal encapsulated SOD against PQ-<strong>induced</strong><br />

human injuries yet.<br />

Vitamin E (α-Tocopherol)<br />

Vitamin E is a lipid-soluble vitamin that exerts its antioxidant effects by<br />

scavenging free radicals and stabilizing membranes containing polyunsaturated fatty<br />

acids (Burton, 1994). The role <strong>of</strong> vitamin E in PQ toxicity was demonstrated in several<br />

81


Part I - General Introduction __________________________________________________<br />

studies where deficiency <strong>of</strong> vitamin E potentiated <strong>the</strong> development <strong>of</strong> acute PQ toxicity<br />

in animals. It was shown that vitamin E deficiency shortened and decreased survival,<br />

worsened histologic lung damage in rats (Block, 1979) and significantly reduced <strong>the</strong><br />

LD50 in mice (Bus et al., 1975) exposed to PQ. Moreover, <strong>the</strong> potentiation <strong>of</strong> acute PQ<br />

toxicity by vitamin E deficiency was reversed by administration <strong>of</strong> vitamin E (Block,<br />

1979). Although <strong>the</strong> mechanism(s) by which vitamin E protects against PQ toxicity is<br />

not fully understood, it may be attributed to its antioxidant properties in preventing LPO<br />

or by scavenging O2 .- and thus preventing its toxicity. Although vitamin E confers<br />

protection against PQ-<strong>induced</strong> injuries in vitamin E-deficient animals, normal animals<br />

receive little benefit from additional pharmacologic supplementation with vitamin E. A<br />

study in male rats, <strong>the</strong> i.p administration <strong>of</strong> vitamin E ei<strong>the</strong>r 30 min after i.p. PQ LD50<br />

challenge followed by a second injection 24 hours later, or 2 hours before PQ challenge<br />

followed by a second injection 26 hours later, did not alter <strong>the</strong> acute mortality nor<br />

reduced <strong>the</strong> characteristic pathological lung changes observed at death (Redetzki et al.,<br />

1980). Moreover, investigating <strong>the</strong> extent <strong>of</strong> lipid peroxidation, expressed as serum<br />

malondialdehyde level, in patients with subacute toxic reactions from PQ poisoning, it<br />

was shown that <strong>the</strong> administration <strong>of</strong> vitamin E to humans (100–4,000 mg/day) was<br />

ineffective in protecting against PQ poisoning and did not affect <strong>the</strong> levels <strong>of</strong><br />

malondialdehyde (Yasaka et al., 1986).<br />

The failure <strong>of</strong> vitamin E to protect against PQ and o<strong>the</strong>r oxidants is unclear at <strong>the</strong><br />

present time. It has been suggested that this ineffectiveness might be related to <strong>the</strong><br />

solubility <strong>of</strong> vitamin E, since lipid-soluble antioxidants take too long to diffuse through<br />

cellular membranes. To overcome this major limitation in patients requiring emergency<br />

treatment, water-soluble analogs <strong>of</strong> α-tocopherol, which can be safely administrated by<br />

i.v. (Petty et al., 1990), or liposomal α-tocopherol preparations (Suntres and Shek,<br />

1995) might <strong>of</strong>fer a better treatment effect. Shahar et al. (Shahar et al., 1980) reported<br />

recovery in a child who had ingested a potentially lethal dose <strong>of</strong> PQ and was treated<br />

with vitamin E. However, Harley et al. (Harley et al., 1977) noted no effect in ano<strong>the</strong>r<br />

case.<br />

82


__________________________________________________Part I - General Introduction<br />

Vitamin C (ascorbic acid)<br />

Ascorbic acid, a water-soluble vitamin, is effective in scavenging free radicals,<br />

including HO . , aqueous peroxyl radicals and O2 .- . Ascorbic acid acts as a two-electron<br />

reducing agent and confers protection by releasing an electron to reduce free radicals,<br />

thus neutralizing <strong>the</strong>se compounds in <strong>the</strong> extracellular aqueous environment, prior to<br />

<strong>the</strong>ir reaction with biological molecules (Evans and Halliwell, 2001). Moreover, <strong>the</strong><br />

antioxidant potential <strong>of</strong> ascorbic acid is not only attributed to its ability to quench ROS,<br />

but also to its ability to regenerate o<strong>the</strong>r small molecule antioxidants, such as α-<br />

tocopherol, GSH and β-carotene (Evans and Halliwell, 2001). Intravenously-<br />

administered vitamin C shortly prior to PQ challenge protected against tissue damage as<br />

evidenced by a reduction <strong>of</strong> <strong>the</strong> exEth, a reliable index <strong>of</strong> oxidative damage (Kang et<br />

al., 1998). Although prior administration <strong>of</strong> ascorbic acid confers protection against PQ<br />

toxicity, <strong>the</strong> use <strong>of</strong> ascorbic acid in treating PQ-<strong>induced</strong> tissue injuries has resulted in<br />

unfavorable consequences. Apparently, ascorbic acid can accelerate <strong>the</strong> generation <strong>of</strong><br />

HO . by reducing oxidized free transition metal ions [e.g. ferric ion (Fe 3+ )] in <strong>the</strong><br />

aqueous phase (Buettner and Jurkiewicz, 1996; Halliwell, 1996; Carr and Frei, 1999;<br />

Evans and Halliwell, 2001). Results show that, during extensive cellular damage,<br />

transition metals are released <strong>into</strong> <strong>the</strong> aqueous phase (Kohen and Chevion, 1985c;<br />

Halliwell, 1996). Ascorbic acid, given at a time when <strong>the</strong> extensive tissue damage<br />

<strong>induced</strong> by PQ is in progress, aggravates <strong>the</strong> oxidative damage (Kang et al., 1998). The<br />

exacerbation <strong>of</strong> <strong>the</strong> oxidative damage following <strong>the</strong> interaction <strong>of</strong> transition metals with<br />

ascorbic acid during <strong>the</strong> progressive stages <strong>of</strong> <strong>paraquat</strong> toxicity, was significantly<br />

reduced by pretreating <strong>the</strong>se animals with desferoxamine (DFO), a chelator that tightly<br />

binds <strong>the</strong> ferric iron just prior to PQ administration (Kang et al., 1998). In PQ poisoned<br />

patients, ascorbic acid showed to be an important free radical scavenger (Hong et al.,<br />

2002).<br />

Desferoxamine (desferrioxamine)<br />

Iron and PQ do appear to behave synergistically in <strong>the</strong> generation <strong>of</strong> HO . , by an<br />

iron-driven Fenton reaction (Fig. 11). Physiologically, free iron (i.e., iron bound to lowmolecular-weight<br />

chelators) exists predominately in <strong>the</strong> ferric (Fe 3+ ) state and <strong>the</strong><br />

83


Part I - General Introduction __________________________________________________<br />

foregoing reaction does not proceed at a toxicologically significant rate. However,<br />

based on evidence from in vitro studies (Vile and Winterbourn, 1988), it has been<br />

suggested that <strong>the</strong> presence <strong>of</strong> PQ facilitate <strong>the</strong> reduction <strong>of</strong> Fe 3+ to ferrous ion (Fe 2+ ),<br />

thus significantly enhancing <strong>the</strong> rate <strong>of</strong> HO . generation as long as sufficient H2O2 is<br />

available (van Asbeck et al., 1989). The reduction <strong>of</strong> Fe 3+ may be achieved directly by<br />

<strong>the</strong> PQ •+ (Vile and Winterbourn, 1988), or indirectly by <strong>the</strong> O2 .- generated through<br />

reduction <strong>of</strong> O2 by PQ •+ (McCord and Day, 1978) (Fig. 11), or by ascorbic acid<br />

(Buettner and Jurkiewicz, 1996; Halliwell, 1996; Carr and Frei, 1999; Evans and<br />

Halliwell, 2001). The importance <strong>of</strong> iron and o<strong>the</strong>r transition metals in <strong>the</strong> PQ-related<br />

damage, has been demonstrated by both in vitro and in vivo studies, where iron<br />

chelation prevented against PQ toxicity (Kohen and Chevion, 1985b; Kohen and<br />

Chevion, 1985c; Kohen and Chevion, 1985a; van Asbeck et al., 1989; Van der Wal et<br />

al., 1992), a treatment that also depends on <strong>the</strong> lipophilicity <strong>of</strong> <strong>the</strong> chelating agents. The<br />

administration <strong>of</strong> DFO by continuous i.v. infusion to vitamin E-deficient rats<br />

significantly reduced mortality produced by PQ (van Asbeck et al., 1989). It has been<br />

shown that DFO can exert its protective effects, not only by inhibiting <strong>the</strong> PQ-<strong>induced</strong><br />

generation <strong>of</strong> HO . , but also by blocking <strong>the</strong> uptake <strong>of</strong> PQ by <strong>the</strong> alveolar type II cells<br />

(Van der Wal et al., 1992). Administration <strong>of</strong> more lipophilic chelating agents, such as<br />

hydroxypyridin-4-one (CP51), also increased <strong>the</strong> survival <strong>of</strong> PQ-challenged rats with a<br />

normal vitamin E status. Moreover, <strong>the</strong> protective effect <strong>of</strong> CP51 was also demonstrated<br />

in vitro experiments where CP51 prevented <strong>the</strong> PQ-<strong>induced</strong> lysis <strong>of</strong> alveolar type II<br />

cells (Van der Wal et al., 1992). Although experimentation with iron chelators against<br />

PQ-<strong>induced</strong> toxicity seems promising, <strong>the</strong> potential efficacy and optimal doses <strong>of</strong> <strong>the</strong><br />

iron chelation <strong>the</strong>rapy in human poisoning have nei<strong>the</strong>r been assessed nor ascertained.<br />

Cl<strong>of</strong>ibrate<br />

Cl<strong>of</strong>ibrate increases <strong>the</strong> expression <strong>of</strong> hepatic catalase, which is an antioxidant<br />

enzyme (Goldenberg et al., 1976). Cl<strong>of</strong>ibrate has a protective effect on PQ-<strong>induced</strong><br />

pulmonary toxicity and on mortality, but only when rats are treated before PQ<br />

administration (Frank et al., 1982). However, cl<strong>of</strong>ibrate has no effect on <strong>the</strong> antoxidant<br />

enzymes in <strong>the</strong> lung, and <strong>the</strong>refore its protection against experimental PQ-<strong>induced</strong> lung<br />

84


__________________________________________________Part I - General Introduction<br />

toxicity seems not to be due to an antioxidant effect (Frank et al., 1982). No clinical<br />

studies for this drug have been reported.<br />

Low molecular weight thiol-containing antioxidants<br />

Since compounds containing SH groups are among <strong>the</strong> most important<br />

endogenous antioxidants, <strong>the</strong>ir <strong>the</strong>rapeutic use has been proposed in oxidant lung injury<br />

(Deneke, 2000). GSH is <strong>the</strong> most abundant non-protein SH in living <strong>organ</strong>isms and it<br />

plays a crucial role in intracellular protection against ROS and o<strong>the</strong>r free radicals<br />

(Anderson, 1997). GSH can function as a nucleophile to form conjugates with many<br />

xenobiotic compounds and/or <strong>the</strong>ir metabolites and can also serve as a reductant in <strong>the</strong><br />

metabolism <strong>of</strong> H2O2 and o<strong>the</strong>r <strong>organ</strong>ic hydroperoxides, a reaction catalyzed by GPx<br />

found in cytosols and mitochondria <strong>of</strong> various cells (Anderson, 1997; Deneke, 2000).<br />

Although in vitro studies have shown that alveolar type II cells can be supplemented<br />

with exogenous GSH to protect against PQ-<strong>induced</strong> injury (Hagen et al., 1986), <strong>the</strong><br />

antioxidant effectiveness <strong>of</strong> exogenously administered GSH for <strong>the</strong> treatment <strong>of</strong><br />

pulmonary injuries against PQ or o<strong>the</strong>r oxidants has been hindered by its rapid<br />

hydrolysis in <strong>the</strong> circulation and its inability to cross cell membranes (Smith et al.,<br />

1992).<br />

N-Acetylcysteine (NAC), <strong>the</strong> acetylated variant <strong>of</strong> <strong>the</strong> amino acid L-cysteine, is a<br />

cell-permeable precursor <strong>of</strong> GSH and an excellent source <strong>of</strong> SH groups (Patterson and<br />

Rhoades, 1988). NAC is converted in <strong>the</strong> body <strong>into</strong> metabolites capable <strong>of</strong> stimulating<br />

GSH syn<strong>the</strong>sis, promoting detoxification and acting directly as free radical scavenger<br />

(Kelly, 1998; Deneke, 2000). It has been shown that <strong>the</strong> administration <strong>of</strong> 20 mg/Kg <strong>of</strong><br />

NAC prior to PQ <strong>into</strong>xication, protects against its toxicity in rats, leading to less edema<br />

and cellular infiltration in <strong>the</strong> lung than control animals without NAC pre-treatment<br />

(Wegener et al., 1988). Also, <strong>the</strong> incubation <strong>of</strong> NAC with alveolar type II cells, which<br />

are known to be specific targets <strong>of</strong> PQ toxicity in vivo, enhanced <strong>the</strong> GSH content <strong>of</strong><br />

<strong>the</strong>se cells and consequently prevented <strong>the</strong> PQ-<strong>induced</strong> cytotoxicity (H<strong>of</strong>fer et al., 1996;<br />

Çeçen et al., 2002). In ano<strong>the</strong>r study, <strong>the</strong> administration <strong>of</strong> NAC to PQ-<strong>into</strong>xicated<br />

animals did not affect <strong>the</strong> survival rate, although it delayed <strong>the</strong> PQ-<strong>induced</strong> release <strong>of</strong><br />

chemoattractants for neutrophils in <strong>the</strong> broncheoalveolar lavage fluid and significantly<br />

reduced <strong>the</strong> infiltration <strong>of</strong> inflammatory cells, suggesting that NAC can confer its<br />

85


Part I - General Introduction __________________________________________________<br />

protective effect by delaying inflammation (H<strong>of</strong>fer et al., 1993; H<strong>of</strong>fer et al., 1996). A<br />

more recent study, using also both in vivo and in vitro experiments, demonstrated that<br />

NAC post-treatment in PQ <strong>into</strong>xicated rats can effectively increase <strong>the</strong> survival rate and<br />

abolish <strong>the</strong> PQ-<strong>induced</strong> oxidative stress and inflammatory response (Yeh et al., 2006).<br />

Different NAC dosage and time schedule administration may explain <strong>the</strong> discrepancies.<br />

In vitro exposure <strong>of</strong> human alveolar cells to PQ produced apoptotic cell death, probably<br />

via oxidative stress <strong>mechanisms</strong> and this toxic effect was inhibited by NAC, an effect<br />

attributed to <strong>the</strong> direct scavenging activity mediated by <strong>the</strong> SH group <strong>of</strong> NAC<br />

(Cappelletti et al., 1998). Clinically, <strong>the</strong>re are a few case reports describing <strong>the</strong><br />

successful treatment <strong>of</strong> PQ poisoned patients by NAC (Lheureux et al., 1995; Drault et<br />

al., 1999; Lopez Lago et al., 2002).<br />

The toxicity <strong>of</strong> PQ in mice was significantly decreased by <strong>the</strong> administration <strong>of</strong><br />

thiosulfite or sulfite, which also abolished <strong>the</strong> PQ-<strong>induced</strong> depletion <strong>of</strong> <strong>the</strong> GSH in liver<br />

<strong>induced</strong> by PQ (Yamamoto, 1993). In culture, cystamine, <strong>the</strong> disulphide form <strong>of</strong> <strong>the</strong><br />

naturally occurring SH group, cysteamine, prevented PQ-<strong>induced</strong> Clara cell damage at<br />

low PQ concentrations (Masek and Richards, 1990). In mice, L-cystine protected against<br />

<strong>the</strong> toxicity <strong>of</strong> PQ by maintaining GSH levels in <strong>the</strong> lung cells (Kojima et al., 1992).<br />

Diethylmaleate, a GSH-depleting agent, increases PQ toxicity in rats (Bus et al., 1975).<br />

No clinical studies utilizing thiosulfite or sulfite treatment have been reported.<br />

Metallothionein (MT) is a metal-binding protein <strong>of</strong> low molecular weight,<br />

containing cysteine as one-third <strong>of</strong> its total amino acids (Deneke, 2000). This protein<br />

has been shown to be an efficient scavenger <strong>of</strong> ROS, such as O2 .- and HO . (Miles et al.,<br />

2000). Syn<strong>the</strong>sis <strong>of</strong> MT can be <strong>induced</strong> by essential metals, such as zinc and copper.<br />

Induction <strong>of</strong> metallothionein in <strong>the</strong> lungs <strong>of</strong> mice after zinc administration has shown to<br />

protect against <strong>the</strong> lethality and pulmonary toxicity <strong>of</strong> PQ (Sato et al., 1996). Although<br />

intrapulmonary MT levels are low, <strong>the</strong>y are readily <strong>induced</strong> by s.c. administration <strong>of</strong> PQ<br />

to mice (Bauman et al., 1991). Nakagawa et al. (Bauman et al., 1991; Nakagawa et al.,<br />

1995; Nakagawa et al., 1998) have also found that PQ-<strong>induced</strong> MT syn<strong>the</strong>sis in <strong>the</strong><br />

liver as a consequence <strong>of</strong> ROS production. The protective role <strong>of</strong> MT in PQ toxicity has<br />

also been demonstrated in transgenic mice deficient in MT genes. In <strong>the</strong>se experiments,<br />

it was shown that tissues in MT-null mice were more susceptible to PQ-<strong>induced</strong><br />

oxidative stress than normal mice, as evidenced by increases in LPO (Sato et al., 1996).<br />

Similarly, Lazo et al. (Lazo et al., 1995) showed that embryonic cells derived from MTnull<br />

mice were more susceptible to ROS produced by PQ. A major reason for <strong>the</strong><br />

86


__________________________________________________Part I - General Introduction<br />

increase in <strong>the</strong> susceptibility <strong>of</strong> <strong>the</strong>se tissues to PQ has been attributed to <strong>the</strong> lower basal<br />

levels <strong>of</strong> non-protein SH groups, including MT and GSH, which constitute <strong>the</strong> first line<br />

<strong>of</strong> defence against oxidative stress-<strong>induced</strong> injuries (Deneke, 2000).<br />

Xanthine Oxidase (XO) Inhibitors<br />

Xanthine dehydrogenase (XD) and xanthine oxidase (XO) are two forms <strong>of</strong> <strong>the</strong><br />

same enzyme that differ in <strong>the</strong> electron acceptor used in <strong>the</strong> final step <strong>of</strong> catalysis. In <strong>the</strong><br />

case <strong>of</strong> XD, <strong>the</strong> final electron acceptor is NAD + (dehydrogenase activity), whereas in<br />

<strong>the</strong> case <strong>of</strong> XO <strong>the</strong> final electron acceptor is O2 (oxidase activity). XD is converted to<br />

XO by oxidation <strong>of</strong> cysteine residues (Cys993 and Cys1326 <strong>of</strong> <strong>the</strong> human enzyme) and/or<br />

proteolytic cleavage. Under normal physiologic conditions, XD is <strong>the</strong> predominant form<br />

<strong>of</strong> <strong>the</strong> enzyme found in vivo. XD/XO catalyzes an important physiologic reaction, <strong>the</strong><br />

sequential oxidation <strong>of</strong> hypoxanthine to xanthine and uric acid. Accordingly to some<br />

studies (Kitazawa et al., 1991; Matsubara et al., 1996), PQ mediates <strong>the</strong> electrontransfer<br />

reaction with XD/XO by reduction/reoxidation cycling. PQ takes electrons<br />

away from reduced XD/XO, reducing itself. Consequently, PQ accelerates <strong>the</strong><br />

generation <strong>of</strong> O2 .- via XD/XO system. In rats fed with a tungsten-enriched diet, which<br />

inhibits <strong>the</strong> XD/XO activity by replacing <strong>the</strong> molybdenium ion within <strong>the</strong> enzyme, <strong>the</strong><br />

mortality due to PQ decreased significantly compared with rats fed with a standard diet<br />

(Kitazawa et al., 1991). Pre-treatment with oxypurinol (1000 mg/Kg s.c.) partially<br />

prevented <strong>the</strong> PQ toxicity in rats. In addition, PQ-exposure showed to increase XO lung<br />

activity (Waintrub et al., 1990). The role <strong>of</strong> XD/XO in PQ toxicity was also investigated<br />

using cultured bovine pulmonary artery endo<strong>the</strong>lial cells (Sakai et al., 1995). Tungsten<br />

and allopurinol inhibited <strong>the</strong> increase <strong>of</strong> XO activity and decreased O2 .- release and <strong>the</strong><br />

subsequent formation <strong>of</strong> o<strong>the</strong>r ROS (Sakai et al., 1993; Sakai et al., 1995). The effects<br />

<strong>of</strong> <strong>the</strong>se treatments have not been investigated in human poisonings.<br />

Selenium<br />

Selenium (Se) is an essential trace element, and a large portion <strong>of</strong> body Se is<br />

present in <strong>the</strong> form <strong>of</strong> cellular GPx (Behne and Wolters, 1983). The Se-containing<br />

enzyme GPx plays an important protective role against PQ. This protective effect <strong>of</strong> Se<br />

87


Part I - General Introduction __________________________________________________<br />

has been reported by several authors (Cagen and Gibson, 1977; Omaye et al., 1978;<br />

Burk et al., 1980). Se-dependent GPx is able to reduce, and <strong>the</strong>reby detoxify, both<br />

<strong>organ</strong>ic and in<strong>organ</strong>ic hydroperoxides using GSH as a reducing agent. More recent<br />

studies indicated that <strong>the</strong> Se-containing enzyme GPx is <strong>the</strong> major, if not <strong>the</strong> only<br />

structural form <strong>of</strong> body Se that protects mice against <strong>the</strong> lethal oxidative stress caused<br />

by high levels <strong>of</strong> PQ; it seems less important, however, in protecting mice against <strong>the</strong><br />

moderate oxidative stress by a low level <strong>of</strong> PQ (Cheng et al., 1998). Se-containing<br />

enzyme GPx plays also a critical role in maintaining <strong>the</strong> redox status <strong>of</strong> mice under<br />

acute oxidative stress, and protects against PQ-<strong>induced</strong> oxidative destruction <strong>of</strong> lipids<br />

and protein in vivo (Cheng et al., 1999). Never<strong>the</strong>less, Se was not yet used in <strong>the</strong><br />

treatment <strong>of</strong> human poisonings.<br />

Niacin and Rib<strong>of</strong>lavin<br />

Niacin (500 mg/Kg BW) decreases <strong>the</strong> mortality rate in rats from 75 to 55%<br />

(Brown et al., Science). At least in <strong>the</strong> isolated perfused rat lung, niacin was shown to<br />

protect against PQ-<strong>induced</strong> lung toxicity (Ghazi-Khansari et al., 2005).<br />

Due to stimulate <strong>the</strong> activity <strong>of</strong> Gred, Sehvartsman et al. (Schvartsman et al.,<br />

1984) observed an improvement <strong>of</strong> <strong>the</strong> survival rate after treatment with rib<strong>of</strong>lavin plus<br />

vitamin C in PQ-<strong>into</strong>xicated rats, while rib<strong>of</strong>lavin given alone was without effect. No<br />

human studies using <strong>the</strong>se vitamins have been reported.<br />

Oils and o<strong>the</strong>r fatty acids<br />

The role <strong>of</strong> nutrients in modulating PQ toxicity in experimental animals has also<br />

been investigated, though not as extensively as for antioxidants. It was noted that an<br />

intramuscular injection <strong>of</strong> commercial corn oil, which was used for <strong>the</strong> administration<br />

<strong>of</strong> lipophilic anti-inflammatory agents, reduced <strong>the</strong> lethality <strong>of</strong> a single oral dose <strong>of</strong> PQ<br />

in mice from 70 to 50%. Similarly, <strong>the</strong> injection <strong>of</strong> o<strong>the</strong>r fresh commercial vegetable<br />

oils bearing different ratios <strong>of</strong> unsaturated to saturated fat as well as fish oils (cod liver<br />

and menhaden oils) also reduced PQ lethality (Fritz et al., 1994). The mechanism<br />

underlying <strong>the</strong> protective effect conferred by <strong>the</strong>se oils is not clear, but it does not<br />

88


__________________________________________________Part I - General Introduction<br />

appear to be due to <strong>the</strong>ir vitamin E content or due to alteration in <strong>the</strong> absorption or<br />

distribution <strong>of</strong> PQ (Fritz et al., 1994). On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> loading <strong>of</strong> hepatocytes<br />

with PUFA (α-linolenic acid) underwent LPO to a greater extent and at much lower PQ<br />

concentrations than normal unloaded hepatocytes (Sugihara et al., 1995). It has been<br />

demonstrated that an increase in monosaturated fatty acids or a reduction in<br />

polyunsaturated fatty acids in lipid membranes decreases <strong>the</strong> susceptibility <strong>of</strong><br />

membranes to oxidant attack (Fritz et al., 1994; Sugihara et al., 1995). The effect <strong>of</strong> soy<br />

protein, soy is<strong>of</strong>lavones and saponins on PQ-<strong>induced</strong> oxidative stress was investigated<br />

in rats. Rats were fed on experimental diets containing casein, soy protein, and casein<br />

with soy is<strong>of</strong>lavones and saponins. The diets were supplemented with 0.025% PQ. The<br />

obtained results suggested that an intake <strong>of</strong> soy protein itself, but not soy is<strong>of</strong>lavones<br />

and saponins, reduces PQ-<strong>induced</strong> oxidative stress in rats (Aoki et al., 2002). These<br />

approaches were never tested in humans.<br />

Angiotensin-converting enzyme inhibitors<br />

Recently, inhibitors <strong>the</strong> angiotensin-converting enzyme (ACE; that catalyse <strong>the</strong><br />

conversion <strong>of</strong> angiotensin I to <strong>the</strong> vasoconstrictor peptide, angiotensin II) have been<br />

reported to prevent PQ-toxicity in animal models (Candan and Alagozlu, 2001;<br />

Mohammadi-Karakani et al., 2006). Several physiological roles <strong>of</strong> angiotensin II have<br />

been clarified not only in relation to <strong>the</strong> pathogenesis <strong>of</strong> hypertension but also regarding<br />

<strong>the</strong> stimulation <strong>of</strong> fibroblast proliferation and collagen syn<strong>the</strong>sis (Booz et al., 1993;<br />

Lasky and Ortiz, 2001). Lisinopril decreased <strong>the</strong> amount <strong>of</strong> hydroxyproline in <strong>the</strong> lung<br />

tissue <strong>of</strong> <strong>the</strong> PQ-exposed rats (Mohammadi-Karakani et al., 2006). The antifibrotic<br />

effect <strong>of</strong> lisinopril was shown to be due to inhibition <strong>of</strong> angiotensin II syn<strong>the</strong>sis, which<br />

results in <strong>the</strong> stimulation <strong>of</strong> fibroblast proliferation and collagen syn<strong>the</strong>sis. Also, when<br />

captopril is administered to PQ poisoned rats prevented PQ toxicity by improving <strong>the</strong><br />

disrupted anti-oxidant capacity, lowering LPO and preventing lung tissue fibrosis<br />

(Candan and Alagozlu, 2001). Lisinopril unlike captopril does not contain SH groups in<br />

its structural formula, which may be <strong>the</strong> reason for lisinopril not having any effect on<br />

LPO (Bagchi et al., 1989). More recently, <strong>the</strong> beneficial effect <strong>of</strong> ACE inhibitors in<br />

preventing pulmonary fibrosis as consequence <strong>of</strong> PQ-exposure was also corroborated by<br />

89


Part I - General Introduction __________________________________________________<br />

Ghazi-Khansari et al. (Ghazi-Khansari et al., 2007). Never<strong>the</strong>less, human studies are<br />

not yet available assessing <strong>the</strong> benefit <strong>of</strong> this treatment.<br />

A second approach has been followed to decrease <strong>the</strong> redox cycling <strong>of</strong> PQ.<br />

Methylene blue, for example, competes 100 to 600 times more effectively than PQ for<br />

reduction by three different flavo-containing enzymes; XO, NADH cytochrome c<br />

reductase, and NADPH cytochrome c reductase, resulting in decreased O2 .- production<br />

(Kelner et al., 1988). However, studies <strong>of</strong> this treatment modality for acute PQ<br />

poisoning are lacking.<br />

A third approach has been followed to prevent <strong>the</strong> accumulation <strong>of</strong> PQ in <strong>the</strong><br />

alveolar epi<strong>the</strong>lial cells via <strong>the</strong> PUS. In tissue culture, spermidine uptake by epi<strong>the</strong>lial<br />

type II cells is inhibited by PQ (Rannels et al., 1985; Rannels et al., 1989). In Clara<br />

cells culture, putrescine and spermidine reduce PQ-<strong>induced</strong> damage, indicating that <strong>the</strong>y<br />

compete for <strong>the</strong> same cell surface receptor (Masek and Richards, 1990). This has been<br />

shown to be possible, in vitro. Studies in vivo, however, have not shown any antidotal<br />

effect (Dunbar et al., 1988). Indeed, putrescine infused to rats and achieving a plasma<br />

concentration fourfold that <strong>of</strong> PQ was unable to decrease ei<strong>the</strong>r its accumulation in <strong>the</strong><br />

lungs or its toxic effects (Dunbar et al., 1988). O<strong>the</strong>r substances such as D-propranolol<br />

and imipramine may decrease <strong>the</strong> pulmonary accumulation <strong>of</strong> PQ in vitro (Drew et al.,<br />

1979). However, in vivo studies did not confirm <strong>the</strong>se data and showed no protective<br />

effect <strong>of</strong> <strong>the</strong>se agents (Drew et al., 1979; Bateman, 1987). Chlorpromazine inhibited PQ<br />

uptake and increased its efflux in vitro (Siddik et al., 1979). Unfortunately, in vivo,<br />

chlorpromazine potentiated <strong>the</strong> PQ toxicity by reducing urinary excretion and increasing<br />

pulmonary PQ concentrations simultaneously (Koyama et al., 1987). In humans,<br />

uncontrolled studies showed no positive effect <strong>of</strong> D-propranolol (Fairshter et al., 1979).<br />

The use <strong>of</strong> anti-PQ antigen-binding fragments (Fab) from cleaved Ab to treat poisoning<br />

or some o<strong>the</strong>r PQ-sequestering agents to remove PQ from lung cells was also tested. Ab<br />

from IgG- and IgM-secreting cell lines have been raised in murine hybridomas and<br />

show high selectivity and affinity for PQ (Bowles et al., 1988; Johnston et al., 1988).<br />

PQ-specific Ab inhibit <strong>the</strong> uptake <strong>of</strong> PQ in vitro by type II alveolar cells from <strong>the</strong> rat<br />

and reduce toxicity (Wright et al., 1987; Chen et al., 1994b). After i.v. injection <strong>of</strong> 0.1<br />

mg/Kg PQ, <strong>the</strong> plasma PQ concentration in rats pre-treated with anti-PQ Ab was<br />

increased and <strong>the</strong> amount excreted in <strong>the</strong> urine was significantly decreased compared<br />

90


__________________________________________________Part I - General Introduction<br />

with controls (Nagao et al., 1989). However, although using anti-PQ Ab can<br />

successfully sequester PQ in <strong>the</strong> plasma compartment <strong>of</strong> rats and mice, it could not<br />

prevent PQ from accumulating in tissues, such as <strong>the</strong> lung, nei<strong>the</strong>r favour its release<br />

(Cadot et al., 1985; Nagao et al., 1989). In fact, such in vitro and in vivo studies suggest<br />

that as <strong>the</strong> concentrations <strong>of</strong> PQ in <strong>the</strong> lung are not changed, PQ Ab nei<strong>the</strong>r prevent PQ<br />

uptake by <strong>the</strong> lung nor favour its release. Moreover, it was predicted that a 100- to 200-<br />

g Fab Ab fragment dose would be required for an adult human, an amount beyond<br />

production capabilities (Wright et al., 1987). More recently a single chain Fv (scFv)<br />

fragment specific for PQ was produced from hybridoma cells secreting a PQ-specific<br />

murine monoclonal Ab, <strong>the</strong> aim being to produce a smaller molecule with high affinity<br />

for PQ (Devlin et al., 1995). However, this scFv fragment was expressed in an insoluble<br />

form and only displayed moderate PQ-binding affinity. Therefore, an attempt was made<br />

to produce a soluble scFv fragment and to increase its PQ binding affinity.<br />

Unfortunately, it became clear that <strong>the</strong> supposed pH dependence <strong>of</strong> PQ binding to <strong>the</strong><br />

scFv fragment was due to tightly bound tris(hydroxymethyl)aminomethane (Tris) from<br />

<strong>the</strong> buffer used to purify <strong>the</strong> Ab (Bowles et al., 1997).<br />

In a fourth approach, <strong>the</strong> effects <strong>of</strong> a lung surfactant-stimulating drug, ambroxol,<br />

or <strong>the</strong> administration <strong>of</strong> exogenous surfactant, have been investigated. Observations that<br />

extensive alveolar collapse represents a relatively early morphological phenomenon in<br />

PQ poisoning, coupled with evidence <strong>of</strong> decreased surface-active material in <strong>the</strong> lung<br />

lavage <strong>of</strong> PQ-treated animals (Robertson et al., 1970; Fisher et al., 1975) have prompted<br />

proposals that surfactant depletion, ei<strong>the</strong>r through a direct action <strong>of</strong> PQ on surfactant<br />

syn<strong>the</strong>sis and/or secretion or as a consequence <strong>of</strong> destruction <strong>of</strong> surfactant-producing<br />

alveolar type II cells, may be a significant event in <strong>the</strong> toxic process and in <strong>the</strong><br />

pathophysiology <strong>of</strong> respiratory failure after PQ <strong>into</strong>xication (Robertson et al., 1970;<br />

Silva and Saldiva, 1998). The poisoning <strong>of</strong> rats with PQ results in a surfactant-deficient<br />

state, due to surfactant inhibition by plasma proteins leaking through <strong>the</strong> damaged<br />

alveoli-capillary membrane (So et al., 1998). In addition, PQ causes a distinct reduction<br />

<strong>of</strong> lecithin fraction to 75% leading to collapse <strong>of</strong> <strong>the</strong> alveoli (Malmgvist et al., 1973).<br />

Intratracheal instillation <strong>of</strong> exogenous surfactant almost completely restored gas<br />

exchange to normal (So et al., 1998). Similar results were also observed after<br />

intratracheal instillation <strong>of</strong> surfactant by improved gas exchange, and prevention <strong>of</strong> lung<br />

inflammation, which resulted in less lung damage as a consequence <strong>of</strong> PQ-exposure<br />

91


Part I - General Introduction __________________________________________________<br />

(Chen et al., 2001; Chen et al., 2002a). In <strong>the</strong> study <strong>of</strong> Salmona et al. (Salmona et al.,<br />

1992) ambroxol pre-treatment increased <strong>the</strong> survival rate <strong>of</strong> <strong>the</strong> animals poisoned with<br />

PQ and antagonized <strong>the</strong> reduction <strong>of</strong> total phospholipid content in <strong>the</strong> lung. Ambroxol<br />

protection also significantly reduced <strong>the</strong> animals death rate in ano<strong>the</strong>r study (Pozzi et<br />

al., 1989). However, in <strong>the</strong> study <strong>of</strong> Nemery et al. (Nemery et al., 1992), ambroxol<br />

treatment did not prevent <strong>the</strong> PQ toxicity. The effects <strong>of</strong> <strong>the</strong>se treatments have not yet<br />

been investigated in human poisonings.<br />

Finally, attempts to reduce <strong>the</strong> extent <strong>of</strong> pulmonary inflammation and fibrosis,<br />

including radio<strong>the</strong>rapy and <strong>the</strong> use <strong>of</strong> anti-inflammatory and immunosuppressant agents<br />

such as cyclophosphamide (CP) and steroids, have not provided compelling evidence <strong>of</strong><br />

clinical efficacy (Bateman, 1987). Immunosuppressive treatment for PQ poisoning was<br />

first reported by Malone in 1971 (Malone et al., 1971) and, since <strong>the</strong>n, this paper<br />

quickly stimulated fur<strong>the</strong>r reports (Eddleston et al., 2003). As described above,<br />

inflammation appears to constitute an early response <strong>of</strong> <strong>the</strong> lung to PQ poisoning. It is<br />

well recognized that inflammatory cells generate ROS, including <strong>the</strong> O2 .- , H2O2, <strong>the</strong><br />

HO . , and hypochlorous acid (Lang et al., 2002; Nagata, 2005; Ricciardolo et al., 2006).<br />

In addition, proteolytic enzymes (such as elastase) are also produced and secreted <strong>into</strong><br />

this environment (Gadek et al., 1984). From a biological perspective, this array <strong>of</strong><br />

deleterious species constitutes an efficient defense against microbiological attack.<br />

However, host cells may also be damaged by <strong>the</strong>se chemical species, and <strong>the</strong>re is<br />

growing evidence to suggest that <strong>the</strong> inflammatory response contributes to <strong>the</strong><br />

pathogenic effect in certain toxic or disease states (Lang et al., 2002; Nagata, 2005;<br />

Ricciardolo et al., 2006). This may be <strong>the</strong> case especially <strong>of</strong> PQ-<strong>induced</strong> toxicity, in<br />

which oxidizing species released by stimulated inflammatory cells fur<strong>the</strong>r increase <strong>the</strong><br />

burden on cellular antioxidant defense systems already "stressed" by <strong>the</strong> initiating PQ<br />

redox cycling. Pulse <strong>the</strong>rapy with CP and methylprednisolone (MP) shows promise in<br />

reducing PQ-related mortality. This <strong>the</strong>rapy is thought to work by reducing <strong>the</strong><br />

inflammatory process leading to pulmonary fibrosis. In a single blinded randomized<br />

clinical trial involving 142 PQ-poisoned patients, pulse <strong>the</strong>rapy reduced mortality in<br />

moderate to severe PQ poisoning from 57% to 18% (Lin et al., 1999). Pulse <strong>the</strong>rapy<br />

included 15 g/Kg/day <strong>of</strong> CP in 5% glucose saline 200 mL given for 2 days and 1 g/day<br />

<strong>of</strong> MP in 200 mL 5% glucose saline i.v. infused for 2 hours for 3 days. All patients<br />

received also dexamethasone (DEX) 10 mg i.v., every 8 hours for 14 days after<br />

92


__________________________________________________Part I - General Introduction<br />

admission. The study was criticized for possible bias during data analysis (Buckley,<br />

2001). Addo and Poon-King (Addo et al., 1984) reported a survival rate <strong>of</strong> 75% in a<br />

group <strong>of</strong> patients treated with a combination <strong>of</strong> high-dosage CP (5 mg/Kg/day, i.v.) and<br />

DEX (24 mg/day i.v.) treatments for 14 days, whereas <strong>the</strong> mortality rate in a historical<br />

control group <strong>of</strong> patients not treated with <strong>the</strong>se two drugs was 80%. However, <strong>the</strong><br />

efficacy <strong>of</strong> this treatment cannot be assessed because criteria <strong>of</strong> severity, such as plasma<br />

PQ concentrations, were not evaluated. Also, <strong>the</strong>se patients were treated with routine<br />

measures, such as Fuller’s Earth, activated charcoal, and magnesium citrate to eliminate<br />

PQ from <strong>the</strong> gut, forced diuresis with furosemide, triamterine, and hydrochlorothiazide<br />

and with niacin and vitamin C as well. The same authors subsequently reported <strong>the</strong>ir<br />

experience with fur<strong>the</strong>r 52 patients, presenting 72 patients in total (Addo and Poon-<br />

King, 1986). Again, <strong>the</strong>y reported a much higher survival rate, 72% compared to 32%,<br />

in patients receiving immunosuppressive <strong>the</strong>rapy compared to historical controls treated<br />

with standard <strong>the</strong>rapy. Perriens and colleagues subsequently reported <strong>the</strong>ir experience <strong>of</strong><br />

using <strong>the</strong> Addo regimen <strong>of</strong> immunosuppression in Suriname (Perriens et al., 1992).<br />

Using a prospective study including 47 consecutive patients with PQ poisoning, <strong>the</strong>re<br />

was no difference in mortality and outcome between <strong>the</strong> 14 patients who had received a<br />

standard treatment and <strong>the</strong> 33 patients who had received a high-dose CP and DEX<br />

treatment (Perriens et al., 1992). O<strong>the</strong>r treatments (Lin et al., 1996; Lin et al., 1999)<br />

have indicated that initial pulse <strong>the</strong>rapy <strong>of</strong> CP 15 mg/Kg/day for 2 days and MP 1 g/day<br />

for 3 days simultaneously, followed by DEX 20 mg/day for 14 days, may be effective in<br />

treating patients with moderate to severe PQ poisoning. However, <strong>the</strong> retrospective<br />

analysis <strong>of</strong> <strong>the</strong>se studies showed that <strong>the</strong>y were not based on an intent-to-treat principle,<br />

as patients who died within 3 or 7 days after <strong>into</strong>xication were excluded from <strong>the</strong> final<br />

analysis. Fur<strong>the</strong>rmore, <strong>the</strong> fact that <strong>the</strong>se studies did not measure plasma PQ levels to<br />

assess <strong>the</strong> severity <strong>of</strong> PQ poisoning may weaken <strong>the</strong>ir results. Recently, Lin et al. (Lin<br />

et al., 2006) reported a novel anti-inflammatory method, with an increase <strong>of</strong> <strong>the</strong><br />

patients’ survival rate, by repeated pulse <strong>the</strong>rapy <strong>of</strong> CP (15 mg/Kg/day, i.v., two days)<br />

and MP (1 g/day i.v., two days) with prolonged DEX [5 mg, i.v., every 6 hours until<br />

PaO2 ≥11.5 kPa (80 mm.Hg)] <strong>the</strong>rapy to treat severely PQ-poisoned patients with 50–<br />

90% predictive mortality (Hart et al., 1984). If PaO2 was 3000/m 3 and <strong>the</strong> duration was >2<br />

weeks after initial CP pulse <strong>the</strong>rapy to avoid a severe leukopenia episode. The results<br />

93


Part I - General Introduction __________________________________________________<br />

were similar to those obtained by previous case reports (Chen et al., 2002b; Lin et al.,<br />

2003). Combined repeated MP pulse <strong>the</strong>rapy preceding continuous DEX is known as a<br />

strong anti-inflammatory treatment in clinical practice (McCune et al., 1988; Boumpas<br />

et al., 1992) suppressing O2 .- production by neutrophils and macrophages. Fur<strong>the</strong>rmore,<br />

CP exerts a wide range <strong>of</strong> immunomodulatory effects that influence virtually all<br />

components <strong>of</strong> <strong>the</strong> cellular and humoral immune response and reduce <strong>the</strong> severity <strong>of</strong><br />

inflammation (Fox and McCune, 1994), <strong>the</strong>refore contributing to <strong>the</strong> overall effect. In<br />

addition, CP-<strong>induced</strong> leukopenia 1-2 weeks later may contribute to reduce pulmonary<br />

inflammatory process <strong>of</strong> PQ-poisoned patients (Addo and Poon-King, 1986; Lin et al.,<br />

1996). Hence, <strong>the</strong> efficacy <strong>of</strong> pulse <strong>the</strong>rapy may be due to prevention and/or reduction<br />

<strong>the</strong> PQ-<strong>induced</strong> severe inflammation in lungs.<br />

94<br />

Due to <strong>the</strong> prevention <strong>of</strong> fibroblasts proliferation, radio<strong>the</strong>rapy has been associated<br />

with successful reversal <strong>of</strong> PQ pulmonary damage (Webb et al., 1984) but has not been<br />

successful in preventing fatality in severe PQ poisonings (Bloodworth et al., 1986).<br />

Irradiation <strong>of</strong> <strong>the</strong> lungs was considered in patients who, after surviving <strong>the</strong> acute phase<br />

<strong>of</strong> poisoning with PQ, showed progressive deterioration <strong>of</strong> respiratory function<br />

(Shirahama et al., 1987). Studies <strong>of</strong> single-dose radiation treatment in mice have not<br />

confirmed that radio<strong>the</strong>rapy has any benefit on PQ-<strong>induced</strong> pulmonary injury (Parkins<br />

and Fowler, 1985; Salovsky and Shopova, 1993). Also in nine cases treated by Talbot et<br />

al. (Talbot and Barnes, 1988), radio<strong>the</strong>rapy failed to show a definite benefit.<br />

8.5 New perspectives<br />

Although many treatments have been proposed and attempted empirically based<br />

on <strong>the</strong> pathologic mechanism <strong>of</strong> toxicity, none are supported by convincing clinical<br />

efficacy. Some authors claim success based solely on <strong>the</strong> results achieved in a single<br />

patient. Few controlled trials <strong>of</strong> <strong>the</strong>se interventions have been performed, and results <strong>of</strong><br />

published case series are inconsistent. Major deficits in assessing clinical benefit from<br />

various interventions and <strong>the</strong>ir combinations include <strong>the</strong> lack <strong>of</strong> a uniformly used<br />

prognostic indicator that reliably predicts risk <strong>of</strong> death at an early stage in <strong>the</strong> poisoning<br />

and small numbers <strong>of</strong> patients receiving a particular intervention. In <strong>the</strong> present section<br />

<strong>the</strong> new and promising treatments are discussed.


__________________________________________________Part I - General Introduction<br />

8.5.1 Mechanical ventilation with additional inhalation <strong>of</strong> NO<br />

Over <strong>the</strong> last years, mechanical ventilation with additional inhalation <strong>of</strong> NO, a<br />

gaseous molecule that contains an unpaired electron, has been proposed for <strong>the</strong><br />

treatment <strong>of</strong> ARDS (Troncy et al., 1997; Hart, 1999). NO has a vasodilator effect in <strong>the</strong><br />

lung areas with a high ventilation/perfusion ratio and this effect results in an increase in<br />

<strong>the</strong> PaO2/FiO2 ratio (Gianetti et al., 2002). Given that PQ toxicity is increased by<br />

oxygenation, NO inhalation in human PQ poisoned patients seems to be promising. This<br />

might permit a period <strong>of</strong> survival long enough for total systemic elimination <strong>of</strong> <strong>the</strong><br />

ingested PQ, at which time lung transplantation might be undertaken without <strong>the</strong> risk <strong>of</strong><br />

PQ-<strong>induced</strong> fibrosis developing in <strong>the</strong> grafted lung(s). Designed to evaluate <strong>the</strong> effects<br />

<strong>of</strong> inhaled NO on <strong>the</strong> PQ-<strong>induced</strong> lung injury in rats, <strong>the</strong> study <strong>of</strong> Cho et al. (Cho et al.,<br />

2005) showed that <strong>the</strong> inhalation <strong>of</strong> NO contributed to increase <strong>the</strong> survival rate, and<br />

also helped to reduce <strong>the</strong> LPO and to inhibit <strong>the</strong> pulmonary fibrosis. Awkwardly,<br />

studies performed by Berisha et al. (Berisha et al., 1994) in isolated guinea pig lungs<br />

supported <strong>the</strong> view that NO is a critical intermediary in <strong>the</strong> production <strong>of</strong> oxidant tissue<br />

damage due to PQ, since all signs <strong>of</strong> injury, including increased airway and perfusion<br />

pressures, pulmonary edema, and protein leakage <strong>into</strong> <strong>the</strong> airspaces, were dosedependently<br />

attenuated or totally prevented by selective and competitive inhibitors <strong>of</strong><br />

NOS such as N G -nitro-L-arginine methyl ester or N ω -nitro-L-arginine. The underlying<br />

mechanism is thought to be due to NO rapid reaction with O2 .- to form <strong>the</strong> strong<br />

oxidant peroxynitrite (ONOO - ) (Nemery and van Klaveren, 1995). An alternative<br />

hypo<strong>the</strong>sis was subsequently proposed, based on <strong>the</strong> findings that PQ uses NOS as an<br />

electron source to generate O2 .- at <strong>the</strong> expense <strong>of</strong> NO (i.e., NOS switches from an<br />

oxygenase to a PQ reductase) (Day et al., 1999). The data reported in this last study<br />

supported <strong>the</strong> concept that NOS is, indeed, a PQ diaphorase, and suggests that toxicity<br />

associated with such redox-active compounds involves a loss <strong>of</strong> NO formation, coupled<br />

with increased O2 .- generation. In accordance to a lower NO production and consequent<br />

inhibition <strong>of</strong> NO-<strong>induced</strong> vascular relaxation (Day et al., 1999), high systolic and<br />

diastolic pressure, measured through a ca<strong>the</strong>ter inserted in <strong>the</strong> carotid artery, was<br />

observed in Wistar rats exposed to PQ (35 mg/Kg, i.p.) (Oliveira et al., 2005). The fact<br />

that PQ increases pulmonary artery and airway pressures emphasizes <strong>the</strong> importance <strong>of</strong><br />

NO deficiency in <strong>the</strong> toxicological response and may explain why patients suffering<br />

from PQ poisoning improve when treated with inhaled NO (Koppel et al., 1994;<br />

95


Part I - General Introduction __________________________________________________<br />

Maruyama et al., 1995; Eisenman et al., 1998). No adverse consequences and<br />

tachyphylaxis were observed at <strong>the</strong> concentrations <strong>of</strong> inhaled NO used. Guidelines from<br />

<strong>the</strong> National Institute for Occupational Safety and Health state that a time-weighted<br />

average <strong>of</strong> 25 ppm for NO constitutes a permissible exposure level (Services, 1988).<br />

The use <strong>of</strong> NO in <strong>the</strong> treatment <strong>of</strong> PQ poisonings definitively deserves fur<strong>the</strong>r studies.<br />

8.5.2 Prop<strong>of</strong>ol<br />

Ano<strong>the</strong>r promising treatment comes from <strong>the</strong> studies <strong>of</strong> Ariyana et al. (Ariyama et<br />

al., 2000) and Lugo-Vallin et al. (Lugo-Vallin Ndel et al., 2002), who both observed an<br />

increase <strong>of</strong> <strong>the</strong> median survival time <strong>of</strong> mice and rats <strong>into</strong>xicated with PQ post-treated<br />

with prop<strong>of</strong>ol, mainly due to its recognized scavenging activity (Murphy et al., 1992).<br />

Because <strong>of</strong> this property, prop<strong>of</strong>ol has been proposed for patients in intensive care units<br />

with multi<strong>organ</strong> failure or ARDS (Smith et al., 1994).<br />

96<br />

9. SEQUELAE IN SURVIVORS<br />

Pulmonary lesions following PQ poisoning are believed to be almost invariably<br />

fatal. Most patients who survive do not develop obvious pulmonary fibrosis at any stage<br />

<strong>of</strong> <strong>the</strong> <strong>into</strong>xication or recovery phase. Rarely, patients who develop mild lung disease<br />

and survive can have a residual restrictive lung disease and impaired gas exchange<br />

(Hudson et al., 1991). The renal dysfunction or failure is considered to follow <strong>the</strong><br />

natural course <strong>of</strong> acute tubular necrosis.<br />

10. LUNG APPEARANCE AT AUTOPSY<br />

The appearance at autopsy depends on exposure dose and on <strong>the</strong> survival time.<br />

Following severe <strong>into</strong>xications <strong>the</strong>re is likely to be caustic injuries <strong>of</strong> <strong>the</strong> lips. The<br />

mucous membranes <strong>of</strong> <strong>the</strong> mouth, pharynx, oesophagus, larynx, and <strong>the</strong> upper trachea


__________________________________________________Part I - General Introduction<br />

are intensely congested and may be covered with a yellowish-green epi<strong>the</strong>lial slough.<br />

The gastric mucosa is likely to be congested and may contain small haemorrhages. The<br />

kidneys may be swollen and perhaps ra<strong>the</strong>r pale, and <strong>the</strong> liver is also somewhat pale.<br />

Centrilobular necrosis <strong>of</strong> <strong>the</strong> liver and renal tubular necrosis was reported by several<br />

authors (Situnayake et al., 1987). Lungs are heavy, filling and holding <strong>the</strong> shape <strong>of</strong> <strong>the</strong><br />

thoracic cavity, congested, oedematous but microscopically <strong>the</strong>re is not yet any fibrosis<br />

or epi<strong>the</strong>lial proliferation. There may be early pneumonia and pulmonary haemorrhage<br />

is common in addition to <strong>the</strong> severe edema. Aspiration pneumonitis (100% <strong>of</strong> cases),<br />

and pneumothorax with pneumomediastinum (18.75% <strong>of</strong> cases), were remarkable<br />

autopsy findings in those dying from PQ poisoning (Daisley and Simmons, 1999). In<br />

cases <strong>of</strong> moderate <strong>into</strong>xications and consequently with longer survivals, <strong>the</strong> appearances<br />

at autopsy are different. The changes in <strong>the</strong> mucous membranes <strong>of</strong> <strong>the</strong> oropharynx have<br />

resolved and <strong>the</strong> liver and kidneys are usually <strong>of</strong> normal appearance, at least on nakedeye<br />

examination, although <strong>the</strong>re may be microscopic evidence <strong>of</strong> cellular damage. The<br />

dramatic changes are to be found in <strong>the</strong> lungs, where <strong>the</strong>y form <strong>the</strong> now-classical<br />

picture <strong>of</strong> <strong>paraquat</strong> poisoning. On gross examination <strong>the</strong> lungs are usually <strong>of</strong> reduced<br />

size, with a solid appearance and <strong>of</strong> dark grey colour. Section reveals a firm, obviously<br />

fibrotic structure, which is apparently completely airless (Marrs and Proudfoot, 2003).<br />

Microscopic examination reveals a grossly abnormal tissue with abundant fibrosis, <strong>of</strong>ten<br />

virtually obliterating <strong>the</strong> alveoli. Many plump fibroblasts are to be seen in alveolar walls<br />

and alveolar spaces. Hyaline membranes are common, possibly a result <strong>of</strong> ventilation<br />

with high concentrations <strong>of</strong> O2. In general <strong>the</strong> longer <strong>the</strong> survival time, <strong>the</strong> more marked<br />

is <strong>the</strong> proliferation <strong>of</strong> fibroblasts in <strong>the</strong> alveoli, and <strong>the</strong> more airless <strong>the</strong> lung tissue<br />

(Carson and Carson, 1976). On whole sections, <strong>the</strong>re is obliteration <strong>of</strong> air spaces with<br />

<strong>multiple</strong> areas <strong>of</strong> fibrosis. There may be active proliferation <strong>of</strong> <strong>the</strong> bronchial epi<strong>the</strong>lium,<br />

forming small adenomata within <strong>the</strong> parenchyma. At later stages, <strong>the</strong>re is less<br />

inflammation.<br />

97


Part I - General Introduction __________________________________________________<br />

98


__________________________________________________Part I - General Introduction<br />

11. REVIEW ARTICLE<br />

Paraquat exposure as an etiological factor<br />

<strong>of</strong> Parkinson's disease<br />

Reprinted from Neurotoxicology 27: 1110–1122<br />

Copyright© (2006) with kind permission from Elsevier Science Inc<br />

99


Part I - General Introduction __________________________________________________<br />

100


Review<br />

Paraquat exposure as an etiological factor <strong>of</strong> Parkinson’s disease<br />

R.J. Dinis-Oliveira a, *, F. Remião a , H. Carmo a , J.A. Duarte b ,<br />

A. Sánchez Navarro c , M.L. Bastos a , F. Carvalho a, *<br />

a<br />

REQUIMTE, Department <strong>of</strong> Toxicology, Faculty <strong>of</strong> Pharmacy, University <strong>of</strong> Porto, Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal<br />

b<br />

Department <strong>of</strong> Sport Biology, Faculty <strong>of</strong> Sport Sciences, University <strong>of</strong> Porto, Rua Dr. Plácido Costa, 91, 4200-450 Porto, Portugal<br />

c<br />

Department <strong>of</strong> Pharmacy and Pharmaceutical Technology, Faculty <strong>of</strong> Pharmacy,<br />

University <strong>of</strong> Salamanca, Avda. Campo Charro s/n 37007, Salamanca, Spain<br />

Received 14 February 2006; accepted 9 May 2006<br />

Available online 3 July 2006<br />

Abstract<br />

Parkinson’s disease (PD) is a multifactorial chronic progressive neurodegenerative disease influenced by age, and by genetic and environmental<br />

factors. The role <strong>of</strong> genetic predisposition in PD has been increasingly acknowledged and a number <strong>of</strong> relevant genes have been identified (e.g.,<br />

genes encoding a-synuclein, parkin, and dardarin), while <strong>the</strong> search for environmental factors that influence <strong>the</strong> pathogenesis <strong>of</strong> PD has only<br />

recently begun to escalate. In recent years, <strong>the</strong> investigation on <strong>paraquat</strong> (PQ) toxicity has suggested that this herbicide might be an environmental<br />

factor contributing to this neurodegenerative disorder. Although <strong>the</strong> biochemical mechanism through which PQ causes neurodegeneration in PD is<br />

not yet fully understood, PQ-<strong>induced</strong> lipid peroxidation and consequent cell death <strong>of</strong> dopaminergic neurons can be responsible for <strong>the</strong> onset <strong>of</strong> <strong>the</strong><br />

Parkinsonian syndrome, thus indicating that this herbicide may induce PD or influence its natural course. PQ has also been recently considered as<br />

an eligible candidate for inducing <strong>the</strong> Parkinsonian syndrome in laboratory animals, and can <strong>the</strong>refore constitute an alternative tool in suitable<br />

animal models for <strong>the</strong> study <strong>of</strong> PD. In <strong>the</strong> present review, <strong>the</strong> recent evidences linking PQ exposure with PD development are discussed, with <strong>the</strong><br />

aim <strong>of</strong> encouraging new perspectives and fur<strong>the</strong>r investigation on <strong>the</strong> involvement <strong>of</strong> environmental agents in PD.<br />

# 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Parkinson’s disease; Environmental factors; Paraquat; Neurotoxicity; Animal models<br />

Contents<br />

NeuroToxicology 27 (2006) 1110–1122<br />

1. Introduction . . . ............................................................................. 1111<br />

1.1. The pathology <strong>of</strong> Parkinson’s disease . . ........................................................ 1111<br />

1.2. Etiology <strong>of</strong> Parkinson’s disease . . ............................................................ 1111<br />

2. Paraquat toxicity ............................................................................. 1112<br />

2.1. Paraquat toxicity mechanism ................................................................ 1112<br />

2.2. Recent studies in <strong>the</strong> Central Nervous System .................................................... 1113<br />

3. Paraquat and Parkinson’s disease—proposed <strong>mechanisms</strong> . ................................................ 1113<br />

3.1. Paraquat induces long-lasting dopamine overflow and reduction <strong>of</strong> dopamine syn<strong>the</strong>sis ....................... 1113<br />

3.2. Paraquat inhibits <strong>the</strong> complex I <strong>of</strong> <strong>the</strong> mitochondrial electron transport chain . . . ........................... 1114<br />

3.3. Paraquat markedly induces a-synuclein up-regulation and aggregation ................................... 1114<br />

4. Permeability <strong>of</strong> blood-brain barrier to <strong>paraquat</strong> and putative uptake by <strong>the</strong> dopamine transporter . . . .................. 1114<br />

Abbreviations: BBB, blood-brain barrier; CNS, Central Nervous System; DA, dopamine; DOPAC, dihydroxyphenylacetic acid; DTCs, dithiocarbamates; ETC,<br />

electron transport chain; GSH, reduced glutathione; GSSG, oxidized glutathione; H2O2, hydrogen peroxide; HO , hydroxyl radical; HVA, homovanillic acid; LBs,<br />

Lewy bodies; MAO, monoamine oxidase; MB, maneb; MPP + , 1-methyl-4-phenyl-2,3-dihypyridinium ion; MPPP, 1-methyl-4-phenyl-propion-oxypiperedine;<br />

MPTP, 1-methyl-4-phenyl-1,2,3,6 tetrahydropyridine; NMDA, N-methyl-D-aspartate; NO, nitric oxide; NOS, nitric oxide synthase; O 2 , superoxide radical;<br />

ONOO , peroxynitrite anion; PD, Parkinson’s disease; PQ, <strong>paraquat</strong>; RNS, reactive nitrogen species; ROS, reactive oxygen species; SN, substantia nigra; SNpc,<br />

substantia nigra pars compacta; TH, tyrosine hydroxylase<br />

* Corresponding authors. Tel.: +351 222078922; fax: +351 222003977.<br />

E-mail addresses: ricardinis@ff.up.pt, ricardinis@sapo.pt (R.J. Dinis-Oliveira), felixdc@ff.up.pt (F. Carvalho).<br />

0161-813X/$ – see front matter # 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.neuro.2006.05.012


5. The inherent susceptibility <strong>of</strong> dopaminergic neurons contributes to <strong>the</strong> <strong>paraquat</strong>-<strong>induced</strong> damage ..................... 1115<br />

6. Epidemiological studies . ....................................................................... 1115<br />

7. Paraquat as a tool for animal models <strong>of</strong> Parkinson’s disease . .............................................. 1115<br />

8. Two insults are more effective than one: <strong>paraquat</strong> + maneb. . .............................................. 1117<br />

9. Concluding remarks ........................................................................... 1118<br />

Acknowledgement . ........................................................................... 1119<br />

References . . ............................................................................... 1119<br />

1. Introduction<br />

Parkinson’s disease (PD), first described by James Parkinson<br />

in 1817, is a chronic progressive neurodegenerative disease,<br />

affecting at least 1% <strong>of</strong> <strong>the</strong> population over <strong>the</strong> age <strong>of</strong> 55<br />

(Rajput, 1992). It is <strong>the</strong> second most common neurodegenerative<br />

disorder after Alzheimer’s disease, with new 5–24 cases per<br />

100,000 population diagnosed every year (Rajput, 1992).<br />

1.1. The pathology <strong>of</strong> Parkinson’s disease<br />

Fully developed PD comprises motor symptoms such as<br />

resting tremor on one or both sides <strong>of</strong> <strong>the</strong> body, rigidity,<br />

bradykinesia, hypokinesia, and postural reflex impairment<br />

(Marsden, 1994). The pathology <strong>of</strong> PD is not fully<br />

understood. In normal brains <strong>the</strong> number <strong>of</strong> nigral cells is<br />

reduced by 4.7–6% per decade between <strong>the</strong> fifth and <strong>the</strong><br />

ninth decade <strong>of</strong> life (Gibb and Lees, 1991),butthislossisnot<br />

sufficient to cause PD (McGeer et al., 1977). The common<br />

feature <strong>of</strong> PD is <strong>the</strong> degeneration <strong>of</strong> <strong>the</strong> neural connection<br />

between <strong>the</strong> substantia nigra (SN) and <strong>the</strong> striatum (Wooten,<br />

1997), two essential brain regions in maintaining normal<br />

motor function (Fig. 1). The striatum receives its dopaminergic<br />

input from neurons <strong>of</strong> substantia nigra pars compacta<br />

Fig. 1. Schematic diagram showing <strong>the</strong> nigrostriatal dopaminergic pathway. A<br />

cross-section <strong>of</strong> human brain shows <strong>the</strong> caudate and putamen, which constitute<br />

<strong>the</strong> striatum. A section through <strong>the</strong> midbrain shows <strong>the</strong> substantia nigra.<br />

Dopaminergic neurons (in red), whose cell bodies are located in <strong>the</strong> SN, send<br />

projections that terminate and release dopamine in <strong>the</strong> striatum. With <strong>the</strong><br />

degeneration <strong>of</strong> <strong>the</strong> dopaminergic pathway, <strong>the</strong>re is a progressive drop in<br />

dopamine release <strong>into</strong> <strong>the</strong> striatum. Striatal dopamine deficiency, in turn, results<br />

in complex changes in <strong>the</strong> brain’s motor circuitry and causes <strong>the</strong> motor deficits<br />

characteristic <strong>of</strong> Parkinson’s disease (for interpretation <strong>of</strong> <strong>the</strong> references to color<br />

in this figure legend, <strong>the</strong> reader is referred to <strong>the</strong> web version <strong>of</strong> <strong>the</strong> article).<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1111<br />

(SNpc) via <strong>the</strong> nigrostriatal pathway (Moore et al., 1971).<br />

Progressive degeneration <strong>of</strong> <strong>the</strong> nigrostriatal dopaminergic<br />

pathway results in pr<strong>of</strong>ound striatal dopamine (DA)<br />

deficiency (Albin et al., 1989; Crossman, 1989; DeLong,<br />

1990; Greenamyre, 1993; Klockge<strong>the</strong>r and Turski, 1989). By<br />

<strong>the</strong> time that <strong>the</strong> clinical manifestations <strong>of</strong> PD are fully<br />

developed, a large proportion (80%) <strong>of</strong> dopaminergic<br />

neurons in <strong>the</strong> SN are already lost, resulting in reduced<br />

syn<strong>the</strong>sis and release <strong>of</strong> DA from <strong>the</strong> striatal nerve terminals<br />

(Lang and Lozano, 1998).<br />

Besides <strong>the</strong> loss <strong>of</strong> SN neurons, ano<strong>the</strong>r important<br />

pathological feature <strong>of</strong> PD is <strong>the</strong> presence <strong>of</strong> neuronal<br />

cytoplasmatic inclusions known as Lewy bodies (LBs) (Gibb<br />

and Lees, 1988; Marsden, 1994) in some surviving nigral<br />

dopaminergic neurons. In PD, LBs are present in <strong>the</strong><br />

dopaminergic neurons <strong>of</strong> SN, as well as in o<strong>the</strong>r brain regions<br />

such as <strong>the</strong> cortex and magnocellular basal forebrain nuclei<br />

(Braak et al., 1995). A major component <strong>of</strong> <strong>the</strong> LBs is <strong>the</strong> asynuclein<br />

protein (thi<strong>of</strong>lavin S-positive staining), and LBs seem<br />

to derive from a-synuclein aggregation (Spillantini et al., 1997,<br />

1998; Uversky, 2003). However, several o<strong>the</strong>r clinical<br />

syndromes are also associated with intracellular a-synuclein<br />

inclusions (synucleinopathies) (Mukaetova-Ladinska and<br />

McKeith, 2006).<br />

1.2. Etiology <strong>of</strong> Parkinson’s disease<br />

Considerable evidence suggests a multifactorial etiology for<br />

PD, involving genetic and environmental factors. The<br />

contribution <strong>of</strong> genetic predisposition to PD has been<br />

investigated in twin studies (Piccini et al., 1999), case-control<br />

studies (Gasser, 1998, 2001; Sveinbjornsdottir et al., 2000), and<br />

in studies identifying mutations in genes encoding a-synuclein<br />

(Kruger et al., 1998; Polymeropoulos et al., 1997; Zarranz et al.,<br />

2004), parkin (Kitada et al., 1998), PINK1 (Valente et al.,<br />

2004), dardarin (Hernandez et al., 2005) and DJ-1 (Bonifati<br />

et al., 2003).<br />

However, inheritance cannot fully explain all PD cases. In<br />

fact, a comprehensive study <strong>of</strong> over 19,000 white male twins<br />

showed that inheritance is not <strong>the</strong> cause <strong>of</strong> sporadic PD (Tanner<br />

et al., 1999). In addition, a-synuclein is found in all LBs, even<br />

in <strong>the</strong> majority <strong>of</strong> <strong>the</strong> idiopathic PD cases without a-synuclein<br />

mutations (Spillantini et al., 1997), thus indicating that<br />

additional <strong>mechanisms</strong> may lead to conformational changes<br />

and consequent protein aggregation.<br />

Numerous environmental risk factors have been associated<br />

with <strong>the</strong> PD as causative agents, ei<strong>the</strong>r in <strong>the</strong> modulation <strong>of</strong> <strong>the</strong>


1112<br />

disease onset and/or on its progression (Di Monte, 2001, 2003;<br />

Di Monte et al., 2002; McCormack et al., 2002; Tanner, 1989;<br />

Tanner and Ben-Shlomo, 1999). Several environmental agents<br />

are known to cause nigrostriatal damage, and may thus<br />

contribute to PD, namely: (i) metals (Altschuler, 1999; Good<br />

et al., 1992; Gorell et al., 1999; Hellenbrand et al., 1996; Hirsch<br />

et al., 1991; Tanner, 1989; Yasui et al., 1992), (ii) solvents<br />

(Davis and Adair, 1999; Hageman et al., 1999; Pezzoli et al.,<br />

1996; Seidler et al., 1996; Uitti et al., 1994), and (iii) carbon<br />

monoxide (Klawans et al., 1982). Additionally, data from<br />

epidemiological studies point to an association between<br />

increased PD risk and specific environmental factors such as<br />

rural residence (Liou et al., 1997; Marder et al., 1998; Morano<br />

et al., 1994), farming (Fall et al., 1999; Gorell et al., 1998; Liou<br />

et al., 1997; Semchuk et al., 1992), drinking water from wells<br />

(Marder et al., 1998; Morano et al., 1994), and exposure to<br />

agricultural chemicals, including <strong>paraquat</strong> (PQ) (Fall et al.,<br />

1999; Gorell et al., 1998; Liou et al., 1997; Semchuk et al.,<br />

1992, 1993; Vanacore et al., 2002).<br />

Given <strong>the</strong> public health implications concerning risk factors<br />

for <strong>the</strong> development <strong>of</strong> PD, <strong>the</strong> study <strong>of</strong> <strong>the</strong> environmental<br />

factors involved in <strong>the</strong> etiology <strong>of</strong> PD has gained renewed<br />

interest <strong>of</strong> <strong>the</strong> scientific and medical community as well as <strong>of</strong><br />

<strong>the</strong> regulatory governmental agencies. In <strong>the</strong> present review <strong>the</strong><br />

recent evidence from epidemiological, clinical, and experimental<br />

work linking <strong>the</strong> widely used herbicide, PQ, to PD<br />

pathology is discussed.<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

2. Paraquat toxicity<br />

2.1. Paraquat toxicity mechanism<br />

The cellular toxicity <strong>of</strong> PQ is essentially due to its redox<br />

cycle (Fig. 2). Paraquat is reduced, mainly by NADPHcytochrome<br />

P-450 reductase (Clejan and Cederbaum, 1989),<br />

NADPH-cytochrome c reductase (Fernandez et al., 1995), and<br />

<strong>the</strong> mitochondrial complex I also known as NADH: ubiquinone<br />

oxidoreductase (Fukushima et al., 1993; Yamada and Fukushima,<br />

1993), to form a PQ monocation free radical (PQ + ). It is<br />

generally accepted that PQ uses cellular diaphorases, which are<br />

a class <strong>of</strong> enzymes that transfer electrons from NAD(P)H to<br />

small molecules, such as PQ (Aziz et al., 1994; Day et al., 1999;<br />

Dicker and Cederbaum, 1991; Liochev and Fridovich, 1994).<br />

The PQ monocation free radical is <strong>the</strong>n rapidly reoxidized in<br />

<strong>the</strong> presence <strong>of</strong> oxygen generating <strong>the</strong> superoxide radical<br />

(O2 )(Busch et al., 1998; Dicker and Cederbaum, 1991). This<br />

<strong>the</strong>n sets <strong>of</strong>f <strong>the</strong> well-known cascade <strong>of</strong> reactions leading to <strong>the</strong><br />

generation <strong>of</strong> o<strong>the</strong>r reactive oxygen species (ROS), mainly<br />

hydrogen peroxide (H 2O 2) and hydroxyl radical (HO ) and <strong>the</strong><br />

consequent cellular deleterious effects. Indeed, hydroxyl<br />

radicals (Busch et al., 1998; Youngman and Elstner, 1981)<br />

have been implicated in <strong>the</strong> initiation <strong>of</strong> membrane damage by<br />

lipid peroxidation during <strong>the</strong> exposure to PQ in vitro (Busch<br />

et al., 1998) and in vivo (Burk et al., 1980; Dicker and<br />

Cederbaum, 1991).<br />

Fig. 2. Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. A, cellular diaphorases; SOD, superoxide dismutase; CAT, catalase; GPX, glutathione<br />

peroxidase; Gred, glutathione reductase; PQ 2+ , <strong>paraquat</strong>; PQ + , <strong>paraquat</strong> cation radical; HMP, hexose monophosphate pathway; FR, Fenton reaction; HWR, Haber-<br />

Weiss reaction.


2.2. Recent studies in <strong>the</strong> Central Nervous System<br />

Studies <strong>of</strong> PQ toxicity have recently focused on its Central<br />

Nervous System (CNS) effects. Unlike <strong>the</strong> exposure to high<br />

levels <strong>of</strong> PQ that mainly produces pulmonary toxicity, chronic<br />

low levels, resulting from prolonged exposure to nonpneumotoxic<br />

doses, may produce damage to <strong>the</strong> basal ganglia<br />

and Parkinsonism. Toxic damage to <strong>the</strong> brain has been observed<br />

in patients who died from PQ poisoning (Grant et al., 1980;<br />

Hughes, 1988). Autopsy findings in cases <strong>of</strong> acute PQ<br />

poisoning showed cerebral damage with edema, haemorrhage<br />

and neural death. However, in <strong>the</strong>se studies, <strong>the</strong> possibility that<br />

<strong>the</strong> observed tissue changes occurred ei<strong>the</strong>r post-mortem or as a<br />

consequence <strong>of</strong> anoxia due to respiratory dysfunction could not<br />

be excluded.<br />

3. Paraquat and Parkinson’s disease—proposed<br />

<strong>mechanisms</strong><br />

3.1. Paraquat induces long-lasting dopamine overflow and<br />

reduction <strong>of</strong> dopamine syn<strong>the</strong>sis<br />

The excitotoxicity <strong>induced</strong> by N-methyl-D-aspartate<br />

(NMDA) receptor activation, associated to Ca 2+ penetration<br />

<strong>into</strong> <strong>the</strong> cells by activation <strong>of</strong> non-NMDA receptors, is a central<br />

mechanism <strong>of</strong> neurodegeneration in several neurological<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1113<br />

diseases (Dugan and Choi, 1999). There is also increasing<br />

evidence that <strong>the</strong> excitotoxic injury plays a critical role in<br />

progressive degeneration <strong>of</strong> DA neurons in PD (Beal, 1998). In<br />

vivo studies on <strong>the</strong> <strong>mechanisms</strong> <strong>of</strong> PQ-<strong>induced</strong> toxicity in <strong>the</strong><br />

striatum, indicated that PQ stimulates glutamate efflux<br />

initiating excitotoxicity mediated by reactive nitrogen species<br />

(RNS). After activation <strong>of</strong> nitric oxide synthase (NOS)<br />

containing neurons, evoked depolarization <strong>of</strong> NMDA receptor<br />

channels and Ca 2+ penetration <strong>into</strong> <strong>the</strong> cells occur by activation<br />

<strong>of</strong> non-NMDA receptor channels (Shimizu et al., 2003a). It<br />

was also shown that <strong>the</strong> elevation <strong>of</strong> extracellular glutamate<br />

levels was PQ dose-dependent. This phenomenon was<br />

observed shortly after PQ administration. However, <strong>the</strong><br />

mechanism by which PQ induces glutamate efflux is yet to<br />

be clarified. The influx <strong>of</strong> Ca 2+ <strong>into</strong> <strong>the</strong> cells triggers <strong>the</strong><br />

mobilization <strong>of</strong> Ca 2+ -dependent intracellular processes including<br />

<strong>the</strong> activation <strong>of</strong> neuronal NOS. Nitric oxide ( NO)<br />

produced by NOS diffuses to dopaminergic terminals, where it<br />

is thought to play an important role in excitotoxicity, probably<br />

through <strong>the</strong> formation <strong>of</strong> <strong>the</strong> peroxynitrite anion (ONOO )<br />

upon reaction with O 2<br />

produced by <strong>the</strong> redox-cycle <strong>of</strong> PQ<br />

(Figs. 2 and 3) (LaVoie and Hastings, 1999). Peroxynitrite,<br />

which is a lipid-permeable ion with a wider range <strong>of</strong> chemical<br />

targets than NO, can oxidize proteins, lipids, RNA, and DNA.<br />

It inhibits <strong>the</strong> function <strong>of</strong> manganese superoxide dismutase,<br />

which can lead to increased O2 and ONOO formation.<br />

Fig. 3. Schematic representation <strong>of</strong> <strong>the</strong> events that occur in neuronal mitochondria. PQ can be converted <strong>into</strong> a radical (accepting one electron) from NADH via<br />

complex I in mitochondrial ETC (Fukushima et al., 1993), blocking mitochondrial electron flow (McCormack et al., 2002). By redox cycling with molecular oxygen,<br />

PQ leads to superoxide anions (O 2 ) formation in dopaminergic neurons. O 2 formation and lipid peroxidation inhibits complex I activity (Fukushima et al., 1994),<br />

and consequently affects mitochondrial function. O2 easily reacts with nitric oxide ( NO) to generate peroxynitrite anion (ONOO ), that can also be responsible for<br />

PQ-<strong>induced</strong> neurotoxicity. ONOO is not only an oxidizing agent on its own but also degrades to form hydroxyl radicals, among o<strong>the</strong>r reactive species (Beckman<br />

et al., 1990). This neurotoxic event could cause a continuous and long-lasting overflow <strong>of</strong> dopamine.


1114<br />

Additionally, ONOO is an effective inhibitor <strong>of</strong> enzymes in<br />

<strong>the</strong> mitochondrial respiratory chain, decreasing ATP syn<strong>the</strong>sis.<br />

Secondly, ONOO damages DNA strands and inhibits DNA<br />

ligase, which increases DNA strand breaks (Ebadi and Sharma,<br />

2003). Noteworthy, long-lasting DA release (Shimizu et al.,<br />

2003a) and consequent death <strong>of</strong> <strong>the</strong> dopaminergic neurons was<br />

prevented by treatment with glutamate receptor antagonists, by<br />

a NOS inhibitor and by <strong>the</strong> monoamine oxidase inhibitor Ldeprenyl,<br />

strongly suggesting that chronic exposure to low PQ<br />

doses leads to an increased vulnerability <strong>of</strong> dopaminergic<br />

neurons in <strong>the</strong> nigrostriatal DA system via <strong>the</strong> excitotoxic<br />

pathway (Shimizu et al., 2003b). A study in PC12 cells showed<br />

that <strong>the</strong> rate-limiting enzyme in DA syn<strong>the</strong>sis, tyrosine<br />

hydroxylase (TH), is a selective target for nitration following<br />

exposure to ONOO (Ara et al., 1998). Nitration <strong>of</strong> tyrosine<br />

residues in TH results in loss <strong>of</strong> enzymatic activity (Ara et al.,<br />

1998). In <strong>the</strong> mouse striatum, tyrosine nitration-mediated loss<br />

<strong>of</strong> TH activity parallels <strong>the</strong> decline in DA levels (Ara et al.,<br />

1998). These results indicate that tyrosine nitration induces TH<br />

inactivation and consequent DA syn<strong>the</strong>sis impairment. Thus,<br />

glutamate-<strong>induced</strong> RNS-mediated cytotoxicity plays an<br />

important role in <strong>the</strong> toxic effect <strong>of</strong> PQ on dopaminergic<br />

terminals.<br />

3.2. Paraquat inhibits <strong>the</strong> complex I <strong>of</strong> <strong>the</strong> mitochondrial<br />

electron transport chain<br />

The mitochondrial complex I (located in <strong>the</strong> inner<br />

mitochondrial membrane and protruded <strong>into</strong> <strong>the</strong> matrix) is<br />

<strong>the</strong> first and <strong>the</strong> most complex <strong>of</strong> <strong>the</strong> three energy-transducing<br />

enzyme complexes <strong>of</strong> <strong>the</strong> mitochondrial electron transport<br />

chain (ETC). It is <strong>the</strong> point <strong>of</strong> entry for <strong>the</strong> major fraction <strong>of</strong><br />

electrons that cross <strong>the</strong> respiratory chain. As a component <strong>of</strong> <strong>the</strong><br />

ETC, complex I oxidizes NADH to NAD + and transfers<br />

electrons to ubiquinone. In addition, it translocates protons<br />

from <strong>the</strong> mitochondrial matrix to <strong>the</strong> intermembrane space,<br />

contributing to <strong>the</strong> electrochemical gradient required for ATP<br />

syn<strong>the</strong>sis (Hatefi, 1985). Several authors suggest that PQ direct<br />

cytotoxicity is <strong>the</strong> consequence <strong>of</strong> a mitochondrial dysfunction<br />

(Blaszczynski et al., 1985; Hirai et al., 1985; Thakar and<br />

Hassan, 1988; Tomita, 1991). Once PQ is reduced to its radical<br />

PQ + (accepting one electron from NADH) via complex I in<br />

mitochondrial ETC (Fukushima et al., 1993), <strong>the</strong> consequent<br />

O2 high production rate may inhibit <strong>the</strong> activity <strong>of</strong> <strong>the</strong><br />

complex I (Fukushima et al., 1994) and thus causing<br />

mitochondrial dysfunction. This redox-cycling can also cause<br />

lipid peroxidation <strong>of</strong> <strong>the</strong> mitochondrial inner membrane<br />

(Yamada and Fukushima, 1993), and as a result, <strong>the</strong> target<br />

tissue may be damaged (Fukushima et al., 1994). Tawara et al.<br />

(1996) proposed that <strong>the</strong> involvement <strong>of</strong> PQ in <strong>the</strong> etiology <strong>of</strong><br />

PD is based on <strong>the</strong> significantly lower activity <strong>of</strong> complex I. The<br />

electron flux through complex I regulates <strong>the</strong> mitochondrial<br />

transition pore permeability (a large Ca 2+ -dependent pore in <strong>the</strong><br />

inner mitochondrial membrane). Under pathological conditions,<br />

mitochondria de-energize and depolarize as a consequence<br />

<strong>of</strong> <strong>the</strong> opening <strong>of</strong> <strong>the</strong> transition pore, leading to<br />

apoptotic or necrotic cell death (Greenamyre et al., 2001).<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

Severe defects in complex I activity depress ATP syn<strong>the</strong>sis,<br />

induce mitochondria depolarization, and favour Ca 2+ deregulation.<br />

The combination <strong>of</strong> all <strong>the</strong>se factors may cause <strong>the</strong><br />

early onset and rapid progression <strong>of</strong> neurological diseases such<br />

as PD. On <strong>the</strong> o<strong>the</strong>r hand, subtle abnormalities <strong>of</strong> complex I<br />

might produce milder, late-onset disorders (Greenamyre et al.,<br />

2001). Supporting this hypo<strong>the</strong>sis, non-familiar sporadic PD<br />

has been characterized by a 15–30% reduction <strong>of</strong> complex I<br />

activity (Schapira et al., 1990).<br />

Importantly, dopaminergic neurons are particularly vulnerable<br />

to complex I inhibitors. Complex I activities in rat brain,<br />

lung and liver have all been shown to decrease with time, with a<br />

significant effect observed 2 h after PQ administration. It was<br />

<strong>the</strong>refore concluded that PQ decreases <strong>the</strong> mitochondrial<br />

complex I activity <strong>of</strong> <strong>the</strong> brain at an early stage after PQ<br />

exposure, even before respiratory dysfunction is observed<br />

(Tawara et al., 1996).<br />

3.3. Paraquat markedly induces a-synuclein up-regulation<br />

and aggregation<br />

The abundant presynaptic protein a-synuclein plays an<br />

important role in <strong>the</strong> formation <strong>of</strong> LBs inclusions involved in <strong>the</strong><br />

pathogenesis <strong>of</strong> PD (Masliah et al., 2000). It has been<br />

hypo<strong>the</strong>sized that pathological changes may arise from<br />

interactions <strong>of</strong> a-synuclein with toxic agents, a likely mechanism<br />

through which environmental risk factors could contribute to <strong>the</strong><br />

pathogenesis <strong>of</strong> PD. Supporting this hypo<strong>the</strong>sis, <strong>the</strong> in vitro<br />

incubation <strong>of</strong> recombinant a-synuclein in <strong>the</strong> presence <strong>of</strong> PQ<br />

resulted in increased protein fibrillation (Uversky et al., 2001,<br />

2002) with clear concentration-dependent accelerating effects<br />

(Manning-Bog et al., 2002), probably due to <strong>the</strong> preferential<br />

binding <strong>of</strong> PQ to a partially folded a-synuclein intermediate.<br />

Accordingly, following in vivo PQ administration, a-synucleincontaining<br />

aggregates were observed in <strong>the</strong> rodent SN (Manning-<br />

Bog et al., 2002). This up-regulation followed a consistent<br />

pattern, with higher a-synuclein levels attained 2 days after each<br />

<strong>of</strong> three weekly PQ injections and with protein levels returning to<br />

control values by day 7 after PQ administration (Manning-Bog<br />

et al., 2002). The up-regulation <strong>of</strong> a-synuclein as a consequence<br />

<strong>of</strong> toxicant insult and <strong>the</strong> direct interaction between <strong>the</strong> protein<br />

and environmental agents are potential <strong>mechanisms</strong> leading to<br />

a-synuclein pathology in neurodegenerative disorders (Manning-Bog<br />

et al., 2002).<br />

4. Permeability <strong>of</strong> blood-brain barrier to <strong>paraquat</strong> and<br />

putative uptake by <strong>the</strong> dopamine transporter<br />

Ano<strong>the</strong>r important feature <strong>of</strong> PQ toxicity is related to its<br />

ability to permeate <strong>the</strong> blood-brain barrier (BBB) <strong>into</strong> <strong>the</strong> CNS.<br />

Paraquat is a charged molecule, with a hydrophilic structure,<br />

low partition coefficient and does not readily cross membranes.<br />

Thus, it is unlikely that <strong>the</strong> passive entry <strong>of</strong> PQ across <strong>the</strong> BBB<br />

leads to a significant accumulation <strong>of</strong> <strong>the</strong> compound in <strong>the</strong><br />

brain. In accordance, it was previously shown that <strong>the</strong><br />

structurally related dopaminergic neurotoxin 1-methyl-4phenyl-2,3-dihypyridinium<br />

ion (MPP + ) must be formed


intracerebrally by monoamine oxidase (MAO)-B in glia or nondopaminergic<br />

neurons, since it cannot cross <strong>the</strong> BBB (Shimizu<br />

et al., 2001). Never<strong>the</strong>less, PQ does cross <strong>the</strong> BBB, with<br />

maximal brain levels evident after 24 h, as compared with<br />

30 min in o<strong>the</strong>r tissues, following subcutaneous administration<br />

(Widdowson et al., 1996). In fact, PQ could be measured in <strong>the</strong><br />

CNS after systemic injection in rodents (Corasaniti et al.,<br />

1998). It is well known that PQ-<strong>induced</strong> lung damage is<br />

initiated, at least partially by an energy-dependent accumulation<br />

<strong>into</strong> <strong>the</strong> lung through an uptake system shared by<br />

endogenous polyamines such as putrescine (Smith, 1982).<br />

However, <strong>the</strong> polyamine transporters are not expressed in <strong>the</strong><br />

BBB (Shin et al., 1985). Recent studies suggest <strong>the</strong> involvement<br />

<strong>of</strong> an active uptake system, <strong>the</strong> BBB neutral amino acid<br />

transporter, in <strong>the</strong> transport <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> CNS (McCormack<br />

and Di Monte, 2003; Shimizu et al., 2001), to <strong>the</strong> detriment <strong>of</strong> a<br />

possible dysfunction <strong>of</strong> BBB caused by PQ itself or by PQ + .<br />

Brain accumulation and neurotoxicity <strong>of</strong> PQ in <strong>the</strong> mouse<br />

model was completely prevented by co-administration <strong>of</strong><br />

simple amino acids, such as L-valine and L-phenylalanine, or<br />

levodopa, which are competitive substrates for <strong>the</strong> same BBB<br />

transporter (McCormack and Di Monte, 2003). Taken toge<strong>the</strong>r,<br />

<strong>the</strong>se findings suggest that active uptake across <strong>the</strong> BBB may be<br />

essential in <strong>the</strong> sequence <strong>of</strong> events that leads to toxin-<strong>induced</strong><br />

nigrostriatal damage. Intake <strong>of</strong> specific dietary elements (e.g.,<br />

amino acids) or <strong>the</strong>rapeutic agents (e.g., levodopa) may also<br />

significantly modulate <strong>the</strong> effects <strong>of</strong> environmental xenobiotics<br />

such as PQ, by changing <strong>the</strong>ir rate <strong>of</strong> uptake <strong>into</strong> <strong>the</strong> brain.<br />

Several studies have been performed to explain how PQ is<br />

taken <strong>into</strong> striatal cells. Similarly to <strong>the</strong> polyamine uptake<br />

system in <strong>the</strong> lung (Dinis-Oliveira et al., 2006b), this uptake<br />

seems to be sodium dependent (Shimizu et al., 2001). It was<br />

reported that PQ is rapidly taken up by nerve terminals isolated<br />

from mouse cerebral cortex, where it induces lipid peroxidation<br />

in a concentration-dependent manner in <strong>the</strong> presence <strong>of</strong><br />

NAD(P)H and ferrous iron (Yang and Sun, 1998). Importantly,<br />

PQ also, in a concentration-dependent manner, reduces <strong>the</strong><br />

number <strong>of</strong> dopaminergic neurons in cultured rat <strong>organ</strong>otypic<br />

midbrain slices (Shimizu et al., 2003b). Since this damage is<br />

prevented by GBR-12909 (a selective inhibitor <strong>of</strong> DA<br />

transport), <strong>the</strong> involvement <strong>of</strong> <strong>the</strong> DA transporter in <strong>the</strong> PQ<br />

uptake <strong>into</strong> <strong>the</strong> striatal cells has been proposed. However, <strong>the</strong><br />

transport <strong>of</strong> PQ through <strong>the</strong> DA transporter remains a<br />

controversial issue (Barlow et al., 2003).<br />

5. The inherent susceptibility <strong>of</strong> dopaminergic neurons<br />

contributes to <strong>the</strong> <strong>paraquat</strong>-<strong>induced</strong> damage<br />

Comparing to o<strong>the</strong>r neuronal cell types, dopaminergic cells<br />

are much more sensitive to oxidative injury due to <strong>the</strong><br />

participation <strong>of</strong> DA in harmful oxidative reactions (Fitsanakis<br />

et al., 2002; Graham, 1978). The activity <strong>of</strong> MAO, which is<br />

involved in DA metabolism, produces H 2O 2 as a normal<br />

byproduct. Moreover, autoxidation <strong>of</strong> DA results in <strong>the</strong><br />

formation <strong>of</strong> ROS (Lotharius and O’Malley, 2000). Never<strong>the</strong>less,<br />

<strong>the</strong> toxicological implications <strong>of</strong> <strong>the</strong> inherent vulnerability<br />

<strong>of</strong> <strong>the</strong> nigrostriatal DA system are still not fully<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1115<br />

understood. One critical feature <strong>of</strong> <strong>the</strong> mammalian SN in PD<br />

that may contribute to its susceptibility to ROS injury is <strong>the</strong><br />

depletion <strong>of</strong> reduced glutathione (GSH) with no change in<br />

oxidized glutathione (GSSG) (Sian et al., 1994a). This appears<br />

to be due to <strong>the</strong> efflux <strong>of</strong> GSH mainly out <strong>of</strong> <strong>the</strong> glia promoted<br />

by g-glutamyltranspeptidase, with a possible additional<br />

increased conversion <strong>of</strong> GSH to GSSG (which is itself<br />

transported out <strong>of</strong> <strong>the</strong> cells by g-glutamyltranspeptidase), in<br />

response to increased intracellular levels <strong>of</strong> H2O2 (Sian et al.,<br />

1994b). Whe<strong>the</strong>r <strong>the</strong> lower level <strong>of</strong> GSH is a cause or a<br />

consequence <strong>of</strong> <strong>the</strong> pathogenic sequence leading to PD remains<br />

to be determined. Although speculative, <strong>the</strong> hypo<strong>the</strong>sis that<br />

chronic low level PQ-<strong>induced</strong> redox cycling within <strong>the</strong> SN<br />

dopaminergic system may exceed <strong>the</strong> oxidative defenses <strong>of</strong><br />

<strong>the</strong>se cells and thus produce deleterious intracellular events,<br />

including <strong>the</strong> activation <strong>of</strong> <strong>the</strong> apoptotic cascade, remains a<br />

likely explanation for <strong>the</strong> association between PQ and PD.<br />

6. Epidemiological studies<br />

A study performed with 120 patients in Taiwan, where <strong>the</strong><br />

herbicide PQ is commonly sprayed over rice fields, showed a<br />

strong association between PQ exposure and PD risk. The<br />

hazard increased by more than six times in individuals who had<br />

been exposed to PQ for more than 20 years (Liou et al., 1997).<br />

These observations were consistent with a dose-dependent<br />

effect and increased with duration <strong>of</strong> pesticide use in<br />

agricultural workers (Liou et al., 1997; Petrovitch et al.,<br />

2002). Occupational PQ exposure in o<strong>the</strong>r 57 cases also showed<br />

association with Parkinsonism in British Columbia (Hertzman<br />

et al., 1990). A door-to-door survey conducted in Taiwan to<br />

estimate <strong>the</strong> PD prevalence and incidence, indicated that <strong>the</strong><br />

environmental factors may be more important than racial<br />

factors in <strong>the</strong> pathogenesis <strong>of</strong> PD (Chen et al., 2001). In a<br />

population-based case-control study in Calgary previous<br />

occupational herbicide use was <strong>the</strong> only significant predictor<br />

<strong>of</strong> PD risk after multivariate statistical analysis (Semchuk et al.,<br />

1992). Seidler et al. (1996) reported a significant association<br />

between PD and pesticide use but not between PD and o<strong>the</strong>r<br />

rural factors in Germany.<br />

When considering environmental pesticide exposure, concern<br />

must be raised upon <strong>the</strong> effects <strong>of</strong> prolonged exposure to<br />

low levels <strong>of</strong> compounds, combination <strong>of</strong> agrochemicals, and<br />

<strong>the</strong> subtle cellular disruptions that may enhance <strong>the</strong> risk for<br />

developing major dysfunctions or disease. For example, <strong>the</strong><br />

combined exposure to PQ and maneb (MB) targets <strong>the</strong><br />

nigrostriatal DA system and induces locomotor impairment<br />

suggesting that this combination may be considered as a<br />

potential environmental risk factor for Parkinsonism (Thiruchelvam<br />

et al., 2000a,b, 2002).<br />

7. Paraquat as a tool for animal models <strong>of</strong> Parkinson’s<br />

disease<br />

Animal models are an invaluable tool for studying <strong>the</strong><br />

pathogenesis and <strong>the</strong>rapeutic intervention strategies <strong>of</strong> human<br />

disease, including PD and in particular, toxicant-<strong>induced</strong>


1116<br />

Parkinsonism. Since PD does not develop spontaneously in<br />

animals, characteristic functional changes have to be mimicked<br />

by neurotoxic agents. Although an ideal model should<br />

reproduce <strong>the</strong> characteristic clinical and pathological features<br />

<strong>of</strong> PD (i.e., animals should develop progressive loss <strong>of</strong><br />

dopaminergic neurons, show deposition <strong>of</strong> LB-like inclusions<br />

in <strong>the</strong> brain and some features <strong>of</strong> L-dopa-responsive movement<br />

disorder), up to now, no animal model was able to entirely<br />

reproduce all <strong>the</strong> features <strong>of</strong> <strong>the</strong> human disease. Albeit, <strong>the</strong>se<br />

models are vital in <strong>the</strong> dissection <strong>of</strong> <strong>the</strong> many different<br />

molecular and biochemical pathways that are combined in <strong>the</strong><br />

final clinical and pathological manifestations <strong>of</strong> PD (Orth and<br />

Tabrizi, 2003).<br />

Presently, <strong>the</strong> MPTP model represents <strong>the</strong> best characterized<br />

PD animal model because it fulfils many <strong>of</strong> <strong>the</strong> criteria for <strong>the</strong><br />

ideal model <strong>of</strong> this disease. The development <strong>of</strong> this model was<br />

based on <strong>the</strong> accidental discovery in <strong>the</strong> early 1980s, when a<br />

Parkinsonian syndrome in young drug addicts was linked to<br />

<strong>the</strong>ir unintentional MPTP intravenous self-administration when<br />

injecting 1-methyl-4-phenyl-propion-oxypiperedine (MPPP),<br />

also known as ‘‘syn<strong>the</strong>tic heroin’’, that was contaminated with<br />

MPTP (Davis et al., 1979; Langston and Ballard, 1983).<br />

Subsequent work revealed that MPTP is not toxic on its own,<br />

but that it easily enters <strong>the</strong> brain where it is metabolized in <strong>the</strong><br />

astrocytes by MAO-B <strong>into</strong> <strong>the</strong> active toxin MPP + . This<br />

neurotoxin displaces DA from intracellular vesicles <strong>into</strong> <strong>the</strong><br />

cytoplasm where auto-oxidation occurs leading to cellular<br />

damage (Lotharius and O’Malley, 2000). MPP + is selectively<br />

transported <strong>into</strong> <strong>the</strong> dopaminergic neurons through <strong>the</strong> DA<br />

transporter, accumulates in mitochondria and inhibits complex<br />

I(Greenamyre et al., 2001), thus acting as a selective complex I<br />

mitochondrial ETC toxicant that produces a Parkinsonian<br />

syndrome similar to <strong>the</strong> idiopathic PD in humans (Langston,<br />

1996). In spite <strong>of</strong> <strong>the</strong> similarity between MPTP-<strong>induced</strong> PD in a<br />

number <strong>of</strong> species (mice, cats, and primates) and <strong>the</strong> sporadic<br />

PD in humans with respect to nigrostriatal dopaminergic<br />

degeneration and <strong>the</strong> characteristic behavioural changes<br />

(MPTP causes tremor, rigidity, akinesia, and postural instability,<br />

which are all successfully treated with L-dopa and DA<br />

agonists), some <strong>of</strong> <strong>the</strong> characteristic features <strong>of</strong> <strong>the</strong> disease<br />

differ, such as <strong>the</strong> lack <strong>of</strong> pronounced LB-related pathology<br />

(Forno et al., 1986; Langston et al., 1999). These differences<br />

suggest that <strong>the</strong> full complement <strong>of</strong> clinical (motor and<br />

cognitive) and pathological (nigrostriatal and extra-nigrostriatal)<br />

features <strong>of</strong> PD is unlikely to be mimicked by a single toxic<br />

insult but would ra<strong>the</strong>r involve multifactorial events, such as<br />

exposure to <strong>multiple</strong> toxicants, genetic factors, gene–toxicants<br />

interactions, and age-related effects.<br />

The unfortunate accident linking MPTP exposure and PD<br />

provided a new insight <strong>into</strong> <strong>the</strong> possibility <strong>of</strong> interaction<br />

between o<strong>the</strong>r environmental agents and PD (Langston and<br />

Ballard, 1984; Tanner and Ben-Shlomo, 1999). If MPTP is<br />

capable <strong>of</strong> inducing neurochemical, pathological, and clinical<br />

features that resemble those <strong>of</strong> idiopathic PD (Di Monte et al.,<br />

2002), similar effects might be caused by o<strong>the</strong>r neurotoxicants.<br />

Shortly after <strong>the</strong> discovery <strong>of</strong> <strong>the</strong> neurotoxicity <strong>of</strong> MPTP, <strong>the</strong><br />

potential involvement <strong>of</strong> pesticides in PD pathology became<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

Fig. 4. Chemical structures <strong>of</strong> <strong>paraquat</strong>, MPTP and MPP + .<br />

obvious. Due to <strong>the</strong> close structural similarity between MPP +<br />

and PQ (Fig. 4), this widely used non-selective contact<br />

herbicide emerged as a putative risk factor <strong>of</strong> PD (Di Monte<br />

et al., 1986). Ironically, in <strong>the</strong> 1960s, MPP + itself had been<br />

tested as an herbicide under <strong>the</strong> commercial name <strong>of</strong> cyperquat<br />

(Di Monte, 2001). Several studies show that <strong>the</strong> exposure <strong>of</strong><br />

mice to PQ may be a suitable experimental model to study <strong>the</strong><br />

<strong>mechanisms</strong> involved in PD (Brooks et al., 1999; McCormack<br />

et al., 2002). However, as a candidate SN toxicant, PQ systemic<br />

delivery must produce <strong>the</strong> loss <strong>of</strong> SN dopaminergic neurons and<br />

<strong>the</strong> subsequent neurobehavioral syndrome with depletion <strong>of</strong><br />

DA terminals within <strong>the</strong> striatum. Initial attempts to establish<br />

convincing evidence for <strong>the</strong> direct link between PQ exposure<br />

and PD failed. The first studies showed that striatal DA levels<br />

did not decrease after <strong>the</strong> systemic administration <strong>of</strong> PQ to<br />

animal models (Perry et al., 1986). Recently, this conclusion<br />

has been re-evaluated by using a stereological technique to<br />

count dopaminergic neurons in <strong>the</strong> SN <strong>of</strong> mice (McCormack<br />

et al., 2002). The authors found that systemic subchronic<br />

exposure to PQ induces dopaminergic neurons cell death in<br />

SNpc, as evaluated by <strong>the</strong> stereological counting <strong>of</strong> THimmunoreactive<br />

and Nissl-stained neurons, without significant<br />

depletion <strong>of</strong> striatal DA. Fur<strong>the</strong>rmore, o<strong>the</strong>r investigators<br />

showed that PQ <strong>induced</strong> a neurobehavioral syndrome<br />

characterized by reduced ambulatory activity (Brooks et al.,<br />

1999), thus fulfilling <strong>the</strong> basic criteria for a neurotoxicant<strong>induced</strong><br />

model <strong>of</strong> PD (Brooks et al., 1999; McCormack et al.,<br />

2002). Given <strong>the</strong> propensity <strong>of</strong> PQ to stimulate lipid<br />

peroxidation, Brooks et al. (1999) proposed that <strong>the</strong> oxidative<br />

stress events due to <strong>the</strong> increased production <strong>of</strong> ROS might be<br />

<strong>the</strong> underlying mechanism responsible for SN toxicity. In spite<br />

<strong>of</strong> nigral degeneration (about 20–30% <strong>of</strong> neurons were lost),<br />

this effect was not accompanied by a significant DA depletion


or behavioural changes, a feature that distinguishes this model<br />

from <strong>the</strong> MPTP and rotenone models (McCormack et al.,<br />

2002). Thus, toxicant exposure may decrease <strong>the</strong> number <strong>of</strong><br />

nigral neurons without triggering acute or major functional<br />

consequences. Whe<strong>the</strong>r <strong>the</strong> effects <strong>of</strong> PQ (e.g., DA depletion)<br />

become evident and progress over time or whe<strong>the</strong>r PQ-<strong>induced</strong><br />

injury predisposes to damage from subsequent toxicant<br />

exposure remains to be determined.<br />

Fur<strong>the</strong>rmore, exposure <strong>of</strong> mice to PQ also led to <strong>the</strong><br />

formation <strong>of</strong> intraneuronal aggregates that were evidenced by<br />

anti-a-synuclein antibodies and thi<strong>of</strong>lavin S staining (a dye that<br />

binds to amyloid fibrils) (Manning-Bog et al., 2002), ano<strong>the</strong>r<br />

extremely important PD feature.<br />

8. Two insults are more effective than one:<br />

<strong>paraquat</strong> + maneb<br />

The hypo<strong>the</strong>sis that a combination <strong>of</strong> environmental risk<br />

factors may result in more severe nigrostriatal injury is<br />

supported by several lines <strong>of</strong> experimental evidence. These<br />

observations are <strong>of</strong> special interest, since humans are likely to<br />

be exposed to a complex mixture <strong>of</strong> chemical agents in <strong>the</strong>ir<br />

residential and occupational environments. PQ is a member <strong>of</strong><br />

only one class <strong>of</strong> agricultural chemicals known to have adverse<br />

effects in <strong>the</strong> nigrostriatal DA system. Complex mixtures <strong>of</strong><br />

several pesticides are <strong>of</strong>ten used in overlapping geographical<br />

areas. Such is <strong>the</strong> case <strong>of</strong> <strong>the</strong> simultaneous use <strong>of</strong> PQ and<br />

diethyldithiocarbamates like <strong>the</strong> manganese ethylenebisdithiocarbamate<br />

[maneb (MB), a dithiocarbamate (DTC) fungicide].<br />

In <strong>the</strong> US, <strong>the</strong> heavy use <strong>of</strong> <strong>the</strong>se chemicals along <strong>the</strong> Pacific<br />

Coast, in <strong>the</strong> Nor<strong>the</strong>ast, <strong>the</strong> Plains states, <strong>the</strong> mid-Atlantic, <strong>the</strong><br />

Sou<strong>the</strong>ast states, and also Texas, where <strong>the</strong>se pesticides are<br />

used ei<strong>the</strong>r separately or combined on <strong>the</strong> same crops (e.g.<br />

tomatoes), may be important for <strong>the</strong> etiological basis <strong>of</strong> PD<br />

(Thiruchelvam et al., 2000a). The extensive geographical<br />

overlap <strong>of</strong> <strong>the</strong> PQ + MB applications and PD prevalence<br />

suggests a possible correlation. This possibility has been<br />

corroborated by animal studies. In fact, <strong>the</strong> co-treatment <strong>of</strong><br />

mice with PQ and MB resulted in potentiated neurotoxicity<br />

(Thiruchelvam et al., 2000a,b). The observed effects were,<br />

moreover, highly selective and irreversible for <strong>the</strong> nigrostriatal<br />

DA system, causing a reduction in motor activity and increased<br />

damage <strong>of</strong> both striatal terminals and nigral cell bodies.<br />

Additionally, it was shown that MB and o<strong>the</strong>r DTCs are able to<br />

alter <strong>the</strong> biodisposition <strong>of</strong> DA and PQ, resulting in a prolonged<br />

exposure to <strong>the</strong>se ROS and RNS generating compounds<br />

(Barlow et al., 2003, 2005). Barlow et al. (2003, 2005) showed,<br />

in striatal synaptosomal vesicles, that some DTCs elicit an<br />

increase in DA accumulation, without altering its influx, but<br />

ra<strong>the</strong>r delaying <strong>the</strong> efflux out <strong>of</strong> <strong>the</strong> synaptosomes. The same<br />

DTCs also increased <strong>the</strong> lung and brain tissue content <strong>of</strong><br />

[ 14 C]PQ in vivo. Thus, certain DTCs and o<strong>the</strong>r agents are<br />

capable <strong>of</strong> converting a non-toxic dose <strong>of</strong> PQ and o<strong>the</strong>r<br />

xenobiotics <strong>into</strong> a toxic dose through alterations in toxicokinetics<br />

(Thiruchelvam et al., 2000b). A common mechanism<br />

whereby selective DTCs might alter <strong>the</strong> kinetics <strong>of</strong> both<br />

[ 3 H]DA in vitro in synaptosomes and [ 14 C]PQ and DA in vivo<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1117<br />

seems to be via direct inhibition <strong>of</strong> an efflux transporter that<br />

transports both compounds out <strong>of</strong> <strong>the</strong> cells (Barlow et al., 2003,<br />

2005). Direct action <strong>of</strong> some DTCs on <strong>the</strong> protein involved in<br />

this transport seems likely, given <strong>the</strong> rapid nature <strong>of</strong> <strong>the</strong> effect,<br />

as opposed to a slower mechanism such as altered transcription<br />

or translation. However, <strong>the</strong> identity <strong>of</strong> this efflux transporter is<br />

yet unknown. Efflux transporters are common in humans and<br />

o<strong>the</strong>r species, have wide tissue expression, and diverse<br />

substrates (Taylor, 2002). There are three large families <strong>of</strong><br />

efflux transporters present in <strong>the</strong> brain and o<strong>the</strong>r <strong>organ</strong>s (Taylor,<br />

2002), each having several members: <strong>the</strong> multidrug resistance<br />

transporters (Ambudkar et al., 1999; Holland and Blight, 1999;<br />

Leslie et al., 2001), <strong>the</strong> monocarboxylic acid transporters and<br />

<strong>the</strong> <strong>organ</strong>ic ion transporters. Recently, we demonstrated that <strong>the</strong><br />

induction <strong>of</strong> <strong>the</strong> syn<strong>the</strong>sis de novo <strong>of</strong> membrane P-glycoprotein<br />

by dexamethasone decreases PQ lung accumulation and<br />

consequently its toxicity (Dinis-Oliveira et al., 2006a). On<br />

<strong>the</strong> o<strong>the</strong>r hand, verapamil, a competitive inhibitor <strong>of</strong> this<br />

transporter (Stein, 1997), when given 1 h before dexamethasone<br />

blocked <strong>the</strong>se protective effects, causing an increase<br />

<strong>of</strong> PQ lung concentration and an aggravation in toxicity<br />

(Dinis-Oliveira et al., 2006a). In <strong>the</strong> light <strong>of</strong> our results, we<br />

hypo<strong>the</strong>size that this efflux impairment could similarly be a<br />

consequence <strong>of</strong> P-glycoprotein inhibition by some DTCs. O<strong>the</strong>r<br />

interesting families <strong>of</strong> efflux transporters are <strong>the</strong> <strong>organ</strong>ic cation<br />

transporter family (OCT1–3) and <strong>the</strong> <strong>organ</strong>ic cation/carnitine<br />

family (OCTN1–3). Several transporters in <strong>the</strong>se families have<br />

been shown to efflux DA, MPTP and MPP + , and to be expressed<br />

in neurons (Busch et al., 1998; Wu et al., 1998; Zhang et al.,<br />

1997). Also, <strong>the</strong> fact that PQ is not metabolized in vivo suggests<br />

that it must be removed from cells by an active transport<br />

process.<br />

Given <strong>the</strong> extreme importance <strong>of</strong> coexposure in epidemiological<br />

and toxicological studies <strong>of</strong> PD in human populations<br />

(Thiruchelvam et al., 2000a,b), <strong>the</strong> interactions <strong>of</strong> toxic<br />

chemicals are valuable models for <strong>the</strong> study <strong>of</strong> <strong>the</strong> potential<br />

<strong>mechanisms</strong> by which environmental exposures can cause<br />

PD. In such models, exposure to a single chemical may be<br />

insufficient to induce major effects, whereas <strong>multiple</strong><br />

concurrent exposures, may preclude homeostatic deregulation<br />

with ensuing neuropathological changes by provoking changes<br />

at <strong>multiple</strong> target sites <strong>of</strong> <strong>the</strong> nigrostriatal DA system<br />

(Thiruchelvam et al., 2000a,b).<br />

Accordingly, a model <strong>of</strong> PD in young adult C57BL/6 mice<br />

based on combined exposure to <strong>the</strong>se pesticides was recently<br />

developed (Thiruchelvam et al., 2000a,b). To this end, C57BL/<br />

6 mice were injected intraperitoneally with ei<strong>the</strong>r PQ at a dose<br />

<strong>of</strong> 5 or 10 mg/kg or MB at a dose <strong>of</strong> 15 or 30 mg/kg alone or in<br />

combination once a week for 4 weeks (Thiruchelvam et al.,<br />

2000a,b). Only <strong>the</strong> combined exposure to both chemicals<br />

produced a sustained decrease in motor activity immediately<br />

after <strong>the</strong> injections. Under <strong>the</strong> same conditions, <strong>the</strong> levels <strong>of</strong> DA<br />

and its metabolites [dihydroxyphenylacetic acid (DOPAC) and<br />

homovanillic acid (HVA)] and <strong>the</strong> efficiency <strong>of</strong> DA turnover<br />

(DOPAC/DA) were shown to increase immediately after<br />

injection. Fur<strong>the</strong>rmore, <strong>the</strong> reductions in TH immunoreactivity<br />

(i.e., <strong>the</strong> decrease in <strong>the</strong> number <strong>of</strong> DA neurons), measured 3


1118<br />

days after <strong>the</strong> last injection, were clearly observed only in<br />

animals with combined PQ + MB exposure. Finally, <strong>the</strong><br />

exposure <strong>of</strong> mice to PQ + MB significantly reduced <strong>the</strong>ir<br />

locomotor activity (Thiruchelvam et al., 2000a,b). The finding<br />

that combined PQ + MB exposure targets <strong>the</strong> nigrostriatal DA<br />

system and induces locomotor impairment suggests that this<br />

combination may be considered as potential environmental risk<br />

factor for Parkinsonism (Thiruchelvam et al., 2000a,b).<br />

Additionally, studies using <strong>the</strong> PQ + MB model have shown<br />

a greater vulnerability <strong>of</strong> males to <strong>the</strong> combined treatment,<br />

which is consistent with observations from epidemiologic<br />

studies <strong>of</strong> PD (Wooten et al., 2004). These studies suggest that<br />

<strong>the</strong> greater incidence <strong>of</strong> <strong>the</strong> disease in males observed in<br />

epidemiologic studies may not only be due to a greater<br />

exposure to environmental risk factors such as pesticides but<br />

may also be related to gender-based physiologic differences.<br />

O<strong>the</strong>r studies also indicate that both aging (Thiruchelvam et al.,<br />

2003) and overexpression <strong>of</strong> mutant human a-synuclein<br />

(Thiruchelvam et al., 2004) enhance <strong>the</strong> PD phenotype<br />

produced by PQ + MB. To investigate <strong>the</strong> influence <strong>of</strong> ageing<br />

<strong>the</strong> effects <strong>of</strong> PQ (10 mg/kg) and MB (30 mg/kg) alone and in<br />

combination were examined in C57BL/6 mice aged 6 weeks, 5<br />

or 18 months old, and were evaluated 2 weeks and 3 months<br />

post-treatment (Thiruchelvam et al., 2003). The findings clearly<br />

demonstrated an age-related enhancement <strong>of</strong> sensitivity to<br />

combined PQ + MB. Additionally, some <strong>of</strong> <strong>the</strong> observed effects<br />

were not only permanent but also increased in magnitude over<br />

time. The first indication that ageing enhanced vulnerability<br />

was that <strong>the</strong> total number <strong>of</strong> treatments had to be abbreviated.<br />

Thiruchelvam et al. (2000a,b) in <strong>the</strong>ir initial study using 6week-old<br />

mice, administered 12 PQ + MB treatments, while<br />

with <strong>the</strong> aged mice it was necessary to stop at six injections,<br />

since <strong>the</strong> 18 month PQ + MB treated mice did not recover <strong>the</strong><br />

locomotor activity 24 h post-treatment (Thiruchelvam et al.,<br />

2003), an outcome that has been shown to accurately predict<br />

underlying dopaminergic changes, particularly dopaminergic<br />

cell loss (Thiruchelvam et al., 2000a,b). Both 5- and 18-monthold<br />

mice showed decreased DA 2 weeks after <strong>the</strong> last PQ + MB<br />

treatment that was still evident 3 months after <strong>the</strong> final<br />

treatment. For <strong>the</strong> 5- and 18-month-old mice groups,<br />

progressive reductions in <strong>the</strong> levels <strong>of</strong> DA metabolites and<br />

DA turnover (DOPAC/DA), between <strong>the</strong> second week and third<br />

month after treatment were most pronounced in <strong>the</strong> 18-monthold<br />

mice group injected with PQ + MB.<br />

To evaluate <strong>the</strong> effect <strong>of</strong> <strong>the</strong> combined exposure to PQ + MB<br />

on <strong>the</strong> overexpression <strong>of</strong> mutant human a-synuclein, transgenic<br />

male mice expressing human wild-type a-synuclein (line hwa-<br />

SYN-5) and human doubly-mutated a-synuclein (A53T and<br />

A30P, line hm 2 a-SYN-39) (6–7 months <strong>of</strong> age) were treated<br />

twice a week for 7 weeks, with a saline vehicle or combined PQ<br />

(5 mg/kg) + MB (15 mg/kg). Ten days after <strong>the</strong> last treatment,<br />

only <strong>the</strong> mice expressing <strong>the</strong> human doubly-mutated asynuclein<br />

line exposed to combined PQ + MB showed a<br />

persistent reduction in locomotor activity ( 70% decrease)<br />

compared to <strong>the</strong> saline treatment (Thiruchelvam et al., 2004).<br />

Thus, combined PQ + MB exposure may be a valuable tool<br />

for inducing PD, since it showed to induce selective, age-<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

related, progressive and irreversible nigrostriatal dopaminergic<br />

system neurotoxicity (Thiruchelvam et al., 2003).<br />

9. Concluding remarks<br />

A number <strong>of</strong> clinical and experimental studies have<br />

increased <strong>the</strong> interest in <strong>the</strong> possibility that environmental<br />

chemicals, including PQ, may be related to <strong>the</strong> development <strong>of</strong><br />

PD (Brooks et al., 1999; Corasaniti et al., 1998; Liou et al.,<br />

1996). PQ seems to be one <strong>of</strong> <strong>the</strong> most eligible herbicides that<br />

may contribute for <strong>the</strong> development <strong>of</strong> PD, given that <strong>the</strong><br />

incidence and development <strong>of</strong> <strong>the</strong> disease and <strong>the</strong> extent <strong>of</strong> PQ<br />

exposure strongly correlate. (Brooks et al., 1999; Corasaniti<br />

et al., 1998; Liou et al., 1996, 1997; Morano et al., 1994).<br />

Fur<strong>the</strong>rmore, PQ administered systematically to experimental<br />

animals induces behavioural and biochemical changes that are<br />

compatible with PD symptoms, such as increased rigidity,<br />

akinesia, tremor and decreased DA concentration (Brooks et al.,<br />

1999; Lindquist et al., 1988). Since many human disorders do<br />

not arise spontaneously in animals, characteristic functional<br />

changes have to be mimicked by neurotoxic agents, and thus<br />

PQ can provide a good model to induce PD symptoms in<br />

experimental animals for <strong>the</strong> study <strong>of</strong> <strong>the</strong> pathogenesis and<br />

<strong>the</strong>rapeutic intervention strategies in this neurodegenerative<br />

disease.<br />

Despite <strong>the</strong> suggestive results <strong>of</strong> epidemiological investigations,<br />

some <strong>of</strong> <strong>the</strong> data are equivocal and more detailed<br />

information about <strong>the</strong> association between PQ exposure and<br />

risk for PD is needed (Koller, 1986). Inconsistencies between<br />

<strong>the</strong> results <strong>of</strong> different studies could be explained, at least<br />

partially, by <strong>the</strong> lack <strong>of</strong> biological markers for PD and <strong>the</strong><br />

consequent variability in case definition and diagnostic<br />

accuracy. The development <strong>of</strong> such biomarkers <strong>of</strong> disease<br />

predisposition, occurrence, and progression in vivo (in contrast<br />

to studies <strong>of</strong> biomarkers in post-mortem brain specimens) is<br />

<strong>the</strong>refore critical in PD <strong>research</strong>. In <strong>the</strong> future, collection <strong>of</strong><br />

more precise data about PQ use should ideally be corroborated<br />

by direct-exposure assessments. The effects <strong>of</strong> PQ exposure<br />

may also vary due to genetic differences among individuals. For<br />

example, <strong>the</strong> lack <strong>of</strong> significant DA depletion, even in <strong>the</strong><br />

presence <strong>of</strong> significant nigral cell loss, and <strong>the</strong> increase in TH<br />

activity caused by PQ suggest that toxicant-<strong>induced</strong> nigrostriatal<br />

injury may remain relatively ‘‘silent’’ (McCormack<br />

et al., 2002). Several explanations may account for <strong>the</strong> lack <strong>of</strong><br />

significant DA depletion after PQ treatment, including <strong>the</strong><br />

possibility that, in contrast to o<strong>the</strong>r toxicants, this herbicide<br />

may preferentially target <strong>the</strong> dopaminergic cell bodies ra<strong>the</strong>r<br />

than its terminals or could be <strong>the</strong> result <strong>of</strong> compensatory<br />

<strong>mechanisms</strong> through which enhanced DA syn<strong>the</strong>sis counteracts<br />

<strong>the</strong> effects <strong>of</strong> terminal damage and restore <strong>the</strong> tissue levels <strong>of</strong><br />

<strong>the</strong> neurotransmitter (McCormack et al., 2002). The contribution<br />

<strong>of</strong> additional environmental and/or genetic risk factors may<br />

also be required for this ‘‘subclinical’’ toxic insult to develop<br />

<strong>into</strong> a complete pathological, neurochemical, and behavioural<br />

syndrome.<br />

Fur<strong>the</strong>r studies on <strong>the</strong> association between PQ and PD<br />

should include assessment <strong>of</strong> PQ exposure in humans and


testing <strong>of</strong> long-term effects in animal models. Also, <strong>the</strong> recurrent<br />

association <strong>of</strong> dopaminergic cell injury with specific <strong>mechanisms</strong><br />

<strong>of</strong> neurotoxicity suggests that putative risk factors may be<br />

screened on <strong>the</strong> basis <strong>of</strong> <strong>the</strong>ir effects on mitochondrial complex I<br />

activity, ROS production, and a-synuclein aggregation. Moreover,<br />

it is also clear that interactive <strong>mechanisms</strong>, such as additive/<br />

synergistic/potentiation, probably underlie <strong>the</strong> effects <strong>of</strong><br />

environmental agents on <strong>the</strong> pathogenesis <strong>of</strong> PD. The study <strong>of</strong><br />

<strong>the</strong>se complex events in human populations and <strong>the</strong> development<br />

<strong>of</strong> animal models <strong>of</strong> such toxic interactions is challenging and<br />

essential to <strong>the</strong> elucidation <strong>of</strong> <strong>the</strong> etiology <strong>of</strong> PD.<br />

Finally, as stated by Di Monte (2003) it is important to<br />

emphasize a multidisciplinary approach for future investigations<br />

on <strong>the</strong> role <strong>of</strong> PD environmental risk factors. A crosslinked<br />

validation <strong>of</strong> clinical, epidemiological, and experimental<br />

evidence should lead to <strong>the</strong> formulation <strong>of</strong> tenable pathogenetic<br />

hypo<strong>the</strong>ses, <strong>the</strong> identification <strong>of</strong> specific risk factors, and <strong>the</strong><br />

design <strong>of</strong> effective preventive strategies. Although <strong>the</strong> effect <strong>of</strong><br />

PQ exposure on <strong>the</strong> development <strong>of</strong> PD is still not<br />

comprehensively explained, supporting evidence is accumulating<br />

that this widely used herbicide can penetrate <strong>the</strong> CNS,<br />

producing lethal injury to SNpc neurons and a consequent<br />

neurobehavioral syndrome. The future investigations in this<br />

field can have dramatic implications in public health, as <strong>the</strong>y<br />

may help in PD prevention via <strong>the</strong> elimination or reduction <strong>of</strong><br />

specific exposure risk factors.<br />

Acknowledgement<br />

Ricardo Dinis, acknowledges FCT for his PhD grant (SFRH/<br />

BD/13707/2003).<br />

References<br />

Albin RL, Young AB, Penney JB. The functional anatomy <strong>of</strong> basal ganglia<br />

disorders. Trends Neurosci 1989;12:366–75.<br />

Altschuler E. Aluminum-containing antacids as a cause <strong>of</strong> idiopathic Parkinson’s<br />

disease. Med Hypo<strong>the</strong>ses 1999;53:22–3.<br />

Ambudkar SV, Dey S, Hrycyna CA, Ramachandra M, Pastan I, Gottesman MM.<br />

Biochemical, cellular, and pharmacological aspects <strong>of</strong> <strong>the</strong> multidrug transporter.<br />

Annu Rev Pharmacol Toxicol 1999;39:361–98.<br />

Ara J, Przedborski S, Naini AB, Jackson-Lewis V, Trifiletti RR, Horwitz J, et al.<br />

Inactivation <strong>of</strong> tyrosine hydroxylase by nitration following exposure to<br />

peroxynitrite and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP).<br />

Proc Natl Acad Sci USA 1998;95:7659–63.<br />

Aziz SM, Lipke DW, Olson JW, Gillespie MN. Role <strong>of</strong> ATP and sodium in<br />

polyamine transport in bovine pulmonary artery smooth cells. Biochem<br />

Pharmacol 1994;48:1611–8.<br />

Barlow BK, Thiruchelvam MJ, Bennice L, Cory-Slechta DA, Ballatori N,<br />

Richfield EK. Increased synaptosomal dopamine content and brain concentration<br />

<strong>of</strong> <strong>paraquat</strong> produced by selective dithiocarbamates. J Neurochem<br />

2003;85:1075–86.<br />

Barlow BK, Lee DW, Cory-Slechta DA, Opanashuk LA. Modulation <strong>of</strong><br />

antioxidant defense systems by <strong>the</strong> environmental pesticide maneb in<br />

dopaminergic cells. Neurotoxicology 2005;26:63–75.<br />

Beal MF. Excitotoxicity and nitric oxide in Parkinson’s disease pathogenesis.<br />

Ann Neurol 1998;44:S110–4.<br />

Beckman JS, Beckman TW, Chen J, Marshall PA, Freeman BA. Apparent<br />

hydroxyl radical production by peroxynitrite: implications for endo<strong>the</strong>lial<br />

injury from nitric oxide and superoxide. Proc Natl Acad Sci USA<br />

1990;87:1620–4.<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1119<br />

Blaszczynski M, Litwinska J, Zaborowska D, Bilinski T. The role <strong>of</strong> respiratory<br />

chain in <strong>paraquat</strong> toxicity in yeast. Acta Microbiol Pol 1985;34:243–54.<br />

Bonifati V, Rizzu P, Baren MJv, Schaap O, Breedveld GJ, Krieger E, et al.<br />

Mutations in <strong>the</strong> DJ-1 gene associated with autosomal recessive early-onset<br />

Parkinsonism. Science 2003;299:256–9.<br />

Braak H, Braak E, Yilmazer D, Schultz C, Vos RAd, Jansen EN. Nigral and<br />

extranigral pathology in Parkinson’s disease. J Neural Transm Suppl<br />

1995;46:15–31.<br />

Brooks AI, Chadwick CA, Gelbard HA, Cory-Slechta DA, Feder<strong>of</strong>f HJ.<br />

Paraquat elicited neurobehavioral syndrome caused by dopaminergic neuron<br />

loss. Brain Res 1999;823:1–10.<br />

Burk RF, Lawrence RA, Lane JM. Liver necrosis and lipid peroxidation in <strong>the</strong><br />

rat as result <strong>of</strong> <strong>paraquat</strong> and diquat administration. Effect <strong>of</strong> selenium<br />

deficiency. J Clin Invest 1980;65:1024–31.<br />

Busch AE, Karbach U, Miska D, Gorboulev V, Akhoundova A, Volk C, et al.<br />

Human neurons express <strong>the</strong> polyspecific cation transporter hOCT2, which<br />

translocates monoamine neurotransmitters, amantadine, and memantine.<br />

Mol Pharm 1998;54:342–52.<br />

Chen RC, Chang SF, Su CL, Chen TH, Yen MF, Wu HM, et al. Prevalence,<br />

incidence, and mortality <strong>of</strong> PD: a door-to-door survey in Ilan county,<br />

Taiwan. Neurology 2001;57:1679–86.<br />

Clejan L, Cederbaum AI. Synergistic interaction between NADPH-cytochrome<br />

P-450 reductase, <strong>paraquat</strong> and iron in <strong>the</strong> generation <strong>of</strong> active oxygen<br />

radicals. Biochem Pharmacol 1989;38:1779–86.<br />

Corasaniti MT, Strongoli MC, Rotiroti D, Bagetta G, Nistico G. Paraquat: a<br />

useful tool for <strong>the</strong> in vivo study <strong>of</strong> <strong>mechanisms</strong> <strong>of</strong> neuronal cell death.<br />

Pharmacol Toxicol 1998;83:1–7.<br />

Crossman AR. Neural <strong>mechanisms</strong> in disorders <strong>of</strong> movement. Comp Biochem<br />

Physiol A 1989;93:141–9.<br />

Davis LE, Adair JC. Parkinsonism from methanol poisoning: benefit from<br />

treatment with anti-Parkinson drugs. Mov Disord 1999;14:520–2.<br />

Davis GC, Williams AC, Markey SP, Ebert MH, Caine ED, Reichert CM, et al.<br />

Chronic Parkinsonism secondary to intravenous injection <strong>of</strong> meperidine<br />

analogues. Psychiatry Res 1979;1:249–54.<br />

Day BJ, Patel M, Calavetta L, Chang LY, Stamler JS. A mechanism <strong>of</strong> <strong>paraquat</strong><br />

toxicity involving nitric oxide synthase. Proc Natl Acad Sci USA 1999;96:<br />

12760–5.<br />

DeLong MR. Primate models <strong>of</strong> movement disorders <strong>of</strong> basal ganglia origin.<br />

Trends Neurosci 1990;13:281–5.<br />

Dicker E, Cederbaum AI. NADH-dependent generation <strong>of</strong> reactive oxygen<br />

species by microsomes in <strong>the</strong> presence <strong>of</strong> iron and redox cycling agents.<br />

Biochem Pharmacol 1991;42:529–35.<br />

Di Monte DA. The role <strong>of</strong> environmental agents in Parkinson’s disease. Clin<br />

Neurosci Res 2001;1:419–26.<br />

Di Monte DA. The environment and Parkinson’s disease: is <strong>the</strong> nigrostriatal<br />

system preferentially targeted by neurotoxins? Lancet Neurol 2003;2:531–8.<br />

Di Monte DA, Sandy MS, Ekstrom G, Smith MT. Comparative-studies on <strong>the</strong><br />

<strong>mechanisms</strong> <strong>of</strong> <strong>paraquat</strong> and 1-methyl-4-phenylpyridine (MPP+) cytotoxicity.<br />

Biochem Biophys Res Commun 1986;137:303–9.<br />

Di Monte DA, Lavasani M, Manning-Bog AB. Environmental factors in<br />

Parkinson’s disease. Neurotoxicology 2002;23:487–502.<br />

Dinis-Oliveira RJ, Remião F, Duarte JA, Sanchez Navarro A, Bastos ML,<br />

Carvalho F. P-glycoprotein induction: an antidotal pathway for <strong>paraquat</strong><strong>induced</strong><br />

lung toxicity. Free Radic Biol Med 2006a;41:1213–24.<br />

Dinis-Oliveira RJ, Valle MJdJ, Bastos ML, Carvalho F, Sanchez Navarro A.<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung. Influence <strong>of</strong> sodium depletion.<br />

Xenobiotica 2006b;36:724–37.<br />

Dugan L, Choi DW. Hypoxic-ischemic brain injury and oxidative stress. In:<br />

Siegel GJ, Agran<strong>of</strong>f BW, Albers RW, Fisher SK, Uhler MD, editors. Basic<br />

neurochemistry. Philadelphia, PA: Lippincott-Raven; 1999. p. 712–29.<br />

Ebadi M, Sharma SK. Peroxynitrite and mitochondrial dysfunction in <strong>the</strong><br />

pathogenesis <strong>of</strong> Parkinson’s disease. Antioxid Redox Signal 2003;5:319–35.<br />

Fall PA, Fredrikson M, Axelson O, Granerus AK. Nutritional and occupational<br />

factors influencing <strong>the</strong> risk <strong>of</strong> Parkinson’s disease: a case-control study in<br />

sou<strong>the</strong>astern Sweden. Mov Disord 1999;14:28–37.<br />

Fernandez Y, Subirade I, Anglade F, Periquet A, Mitjavila S. Microsomal<br />

membrane peroxidation by an Fe3+/<strong>paraquat</strong> system. Consequences <strong>of</strong><br />

phenobarbital induction. Biol Trace Elem Res 1995;47:9–15.


1120<br />

Fitsanakis V, Amarnath V, Moore J, Montine K, Zhang J, Montine T. Catalysis<br />

<strong>of</strong> catechol oxidation by metal-dithiocarbamate complexes in pesticides.<br />

Free Radic Biol Med 2002;33:1714–23.<br />

Forno LS, Langston JW, DeLanney LE, Irwin I, Ricaurte GA. Locus ceruleus<br />

lesions and eosinophilic inclusions in MPTP-treated monkeys. Ann Neurol<br />

1986;20:449–55.<br />

Fukushima T, Yamada K, Isobe A, Shiwaku K, Yamane Y. Mechanism <strong>of</strong><br />

cytotoxicity <strong>of</strong> <strong>paraquat</strong>. I. NADH oxidation and <strong>paraquat</strong> radical formation<br />

via complex I. Exp Toxicol Pathol 1993;45:345–9.<br />

Fukushima T, Yamada K, Hojo N, Isobe A, Shiwaku K, Yamane Y. Mechanism<br />

<strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. III. The effects <strong>of</strong> acute <strong>paraquat</strong> exposure on<br />

<strong>the</strong> electron transport system in rat mitochondria. Exp Toxicol Pathol<br />

1994;46:437–41.<br />

Gasser T. Genetics <strong>of</strong> Parkinson’s disease. Ann Neurol 1998;44:S53–7.<br />

Gasser T. Genetics <strong>of</strong> Parkinson’s disease. J Neurol 2001;248:833–40.<br />

Gibb WR, Lees AJ. The relevance <strong>of</strong> <strong>the</strong> Lewy body to <strong>the</strong> pathogenesis <strong>of</strong><br />

idiopathic Parkinson’s disease. J Neurol Neurosurg Psychiatry 1988;51:<br />

745–52.<br />

Gibb WR, Lees AJ. Anatomy, pigmentation, ventral and dorsal subpopulations<br />

<strong>of</strong> <strong>the</strong> substantia nigra, and differential cell death in Parkinson’s disease. J<br />

Neurol Neurosurg Psychiatry 1991;54:388–96.<br />

Good PF, Olanow CW, Perl DP. Neuromelanin-containing neurons <strong>of</strong> <strong>the</strong><br />

substantia nigra accumulate iron and aluminum in Parkinson’s disease: a<br />

LAMMA study. Brain Res 1992;593:343–6.<br />

Gorell JM, Johnson CC, Rybicki BA, Peterson EL, Richardson RJ. The risk <strong>of</strong><br />

Parkinson’s disease with exposure to pesticides, farming, well water, and<br />

rural living. Neurology 1998;50:1346–50.<br />

Gorell JM, Johnson CC, Rybicki BA, Peterson EL, Kortsha GX, Brown GG, et<br />

al. Occupational exposure to manganese, copper, lead, iron, mercury and<br />

zinc and <strong>the</strong> risk <strong>of</strong> Parkinson’s disease. Neurotoxicology 1999;20:239–47.<br />

Graham DG. Oxidative pathways for catecholamines in <strong>the</strong> genesis <strong>of</strong> neuromelanin<br />

and cytotoxic quinones. Mol Pharmacol 1978;14:633–43.<br />

Grant H, Lantos PL, Parkinson C. Cerebral damage in <strong>paraquat</strong> poisoning.<br />

Histopathology 1980;4:185–95.<br />

Greenamyre JT. Glutamate–dopamine interactions in <strong>the</strong> basal ganglia: relationship<br />

to Parkinson’s disease. J Neural Transm Gen Sect 1993;91:255–69.<br />

Greenamyre JT, Sherer TB, Betarbet R, Panov AV, Complex I. Parkinson’s<br />

disease. IUBMB Life 2001;52:135–41.<br />

Hageman G, Hoek Jvd, Hout Mv, Laan Gvd, Steur EJ, Bruin Wd, et al.<br />

Parkinsonism, pyramidal signs, polyneuropathy, and cognitive decline after<br />

long-term occupational solvent exposure. J Neurol 1999;246:198–206.<br />

Hatefi Y. The mitochondrial electron transport and oxidative phosphorylation<br />

system. Annu Rev Biochem 1985;54:1015–69.<br />

Hellenbrand W, Boeing H, Robra BP, Seidler A, Vieregge P, Nischan P, et al.<br />

Parkinson’s disease. II. A possible role for <strong>the</strong> past intake <strong>of</strong> specific<br />

nutrients. Results from a self-administered food-frequency questionnaire<br />

in a case-control study. Neurology 1996;47:644–50.<br />

Hernandez D, Ruiz CP, Crawley A, Malkani R, Werner J, Gwinn-Hardy K, et al.<br />

The dardarin G 2019 S mutation is a common cause <strong>of</strong> Parkinson’s disease<br />

but not o<strong>the</strong>r neurodegenerative diseases. Neurosci Lett 2005;389:137–9.<br />

Hertzman C, Wiens M, Bowering D, Snow B, Calne D. Parkinson’s disease: a<br />

case-control study <strong>of</strong> occupational and environmental risk factors. Am J Ind<br />

Med 1990;17:349–55.<br />

Hirai K, Witschi H, Cote MG. Mitochondrial injury <strong>of</strong> pulmonary alveolar<br />

epi<strong>the</strong>lial cells in acute <strong>paraquat</strong> <strong>into</strong>xication. Exp Mol Pathol 1985;43:242–<br />

52.<br />

Hirsch EC, Brandel JP, Galle P, Javoy-Agid F, Agid Y. Iron and aluminum<br />

increase in <strong>the</strong> substantia nigra <strong>of</strong> patients with Parkinson’s disease: an Xray<br />

microanalysis. J Neurochem 1991;56:446–51.<br />

Holland IB, Blight MA. ABC-ATPases, adaptable energy generators fueling<br />

transmembrane movement <strong>of</strong> a variety <strong>of</strong> molecules in <strong>organ</strong>isms from<br />

bacteria to humans. J Mol Biol 1999;293:381–99.<br />

Hughes JT. Brain damage due to <strong>paraquat</strong> poisoning: a fatal case with<br />

neuropathological examination <strong>of</strong> <strong>the</strong> brain. Neurotoxicology 1988;9:<br />

243–8.<br />

Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y, Minoshima S, et<br />

al. Mutations in <strong>the</strong> parkin gene cause autosomal recessive juvenile<br />

Parkinsonism. Nature 1998;392:605–8.<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

Klawans HJ, Stein RW, Tanner CM, Goetz CG. A pure Parkinsonian syndrome<br />

following acute carbon monoxide <strong>into</strong>xication. Arch Neurol 1982;39:<br />

302–4.<br />

Klockge<strong>the</strong>r T, Turski L. Excitatory amino acids and <strong>the</strong> basal ganglia:<br />

implications for <strong>the</strong> <strong>the</strong>rapy <strong>of</strong> Parkinson’s disease. Trends Neurosci<br />

1989;12:285–6.<br />

Koller WC. Paraquat and Parkinson’s disease. Neurology 1986;36:1147.<br />

Kruger R, Kuhn W, Muller T, Woitalla D, Graeber M, Kosel S, et al. Ala30Pro<br />

mutation in <strong>the</strong> gene encoding alpha-synuclein in Parkinsons disease. Nat<br />

Genet 1998;18:106–8.<br />

Lang AE, Lozano AM. Parkinsons disease-first <strong>of</strong> two parts. N Engl J Med<br />

1998;339:1044–53.<br />

Langston JW. The etiology <strong>of</strong> Parkinsons disease with emphasis on <strong>the</strong> MPTP<br />

story. Neurology 1996;47:S153–60.<br />

Langston JW, Ballard PA. Parkinson’s disease in a chemist working with 1methyl-4-phenyl-1,2,5,6-tetrahydropyridine.<br />

N Engl J Med 1983;309:310.<br />

Langston JW, Ballard P. Parkinsonism <strong>induced</strong> by 1-methyl-4-phenyl-1,2,3,6tetrahydropyridine<br />

(MPTP): implications for treatment and <strong>the</strong> pathogenesis<br />

<strong>of</strong> Parkinson’s disease. Can J Neurol Sci 1984;11:160–5.<br />

Langston JW, Forno LS, Tetrud J, Reeves AG, Kaplan JA, Karluk D. Evidence<br />

<strong>of</strong> active nerve cell degeneration in <strong>the</strong> substantia nigra <strong>of</strong> humans years<br />

after 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine exposure. Ann Neurol<br />

1999;46:598–605.<br />

LaVoie MJ, Hastings TG. Peroxynitrite- and nitrite-<strong>induced</strong> oxidation <strong>of</strong><br />

dopamine: implications for nitric oxide in dopaminergic cell loss. J Neurochem<br />

1999;73:2546–54.<br />

Leslie EM, Deely RG, Cole SP. Toxicological relevance <strong>of</strong> <strong>the</strong> multidrug<br />

resistance protein 1, MRP1 (ABCC1) and related transporters. Toxicology<br />

2001;167:3–23.<br />

Lindquist NG, Larsson BS, Lyden-Sokolowski A. Autoradiography <strong>of</strong><br />

[14C]<strong>paraquat</strong> or [14C]diquat in frogs and mice: accumulation in neuromelanin.<br />

Neurosci Lett 1988;93:1–6.<br />

Liochev SI, Fridovich I. Paraquat diaphorases in Escherichia coli. Free Radic<br />

Biol Med 1994;16:555–9.<br />

Liou HH, Chen RC, Tsai YF, Chen WP, Chang YC, Tsai MC. Effects <strong>of</strong><br />

<strong>paraquat</strong> on <strong>the</strong> substantia nigra <strong>of</strong> <strong>the</strong> wistar rats: neurochemical,<br />

histological, and behavioral studies. Toxicol Appl Pharmacol 1996;137:<br />

34–41.<br />

Liou HH, Tsai MC, Chen CJ, Jeng JS, Chang YC, Chen SY, et al. Environmental<br />

risk factors and Parkinsons disease: a case-control study in Taiwan.<br />

Neurology 1997;48:1583–8.<br />

Lotharius J, O’Malley KL. The parkinsonism-inducing drug 1-methyl-4-phenylpyridinium<br />

triggers intracellular dopamine oxidation. A novel mechanism<br />

<strong>of</strong> toxicity. J Biol Chem 2000;275:38581–8.<br />

Manning-Bog AB, McCormack AL, Li J, Uversky VN, Fink AL, Di Monte DA.<br />

The herbicide <strong>paraquat</strong> causes up-regulation and aggregation <strong>of</strong> alphasynuclein<br />

in mice: <strong>paraquat</strong> and alpha-synuclein. J Biol Chem 2002;277:<br />

1641–4.<br />

Marder K, Logroscino G, Alfaro B, Mejia H, Halim A, Louis E, et al.<br />

Environmental risk factors for Parkinsons disease in an urban multiethnic<br />

community. Neurology 1998;50:279–81.<br />

Marsden CD. Parkinson’s disease. J Neurol Neurosurg Psychiatry 1994;57:<br />

672–81.<br />

Masliah E, Rockenstein E, Veinbergs I, Mallory M, Hashimoto M, Takeda A, et<br />

al. Dopaminergic loss and inclusion body formation in alpha-synuclein<br />

mice: implications for neurodegenerative disorders. Science 2000;287:<br />

1265–9.<br />

McCormack AL, Di Monte DA. Effects <strong>of</strong> L-dopa and o<strong>the</strong>r amino acids<br />

against <strong>paraquat</strong>-<strong>induced</strong> nigrostriatal degeneration. J Neurochem<br />

2003;85:82–6.<br />

McCormack AL, Thiruchelvam M, Manning-Bog AB, Thiffault C, Langston<br />

JW, Cory-Slechta DA, et al. Environmental risk factors and Parkinson’s<br />

disease: selective degeneration <strong>of</strong> nigral dopaminergic neurons caused by<br />

<strong>the</strong> herbicide <strong>paraquat</strong>. Neurobiol Dis 2002;10:119–27.<br />

McGeer PL, McGeer EG, Suzuki JS. Aging and extrapyramidal function. Arch<br />

Neurol 1977;34:33–5.<br />

Moore RY, Bhatnagar RK, Heller A. Anatomical and chemical studies <strong>of</strong> a<br />

nigro-neostriatal projection in <strong>the</strong> cat. Brain Res 1971;30:119–35.


Morano A, Jimenez-Jimenez FJ, Molina JA, Antolin MA. Risk-factors for<br />

Parkinsons disease: case-control study in <strong>the</strong> province <strong>of</strong> Caceres, Spain.<br />

Acta Neurol Scand 1994;89:164–70.<br />

Mukaetova-Ladinska EB, McKeith IG. Pathophysiology <strong>of</strong> synuclein aggregation<br />

in Lewy body disease. Mech Ageing Dev 2006;127:188–202.<br />

Orth M, Tabrizi SJ. Models <strong>of</strong> Parkinsons disease. Mov Disord 2003;18:729–37.<br />

Perry TL, Yong VW, Wall RA, Jones K. Paraquat and two endogenous<br />

analogues <strong>of</strong> <strong>the</strong> neurotoxic substance N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine<br />

do not damage dopaminergic nigrostriatal neurons in <strong>the</strong><br />

mouse. Neurosci Lett 1986;69:285–9.<br />

Petrovitch H, Ross GW, Abbot RD, Sanderson WT, Sharp DS, Tanner CM, et<br />

al. Plantation work and risk <strong>of</strong> Parkinson’s disease in a population-based<br />

longitudinal study. Arch Neurol 2002;59:1787–92.<br />

Pezzoli G, Strada O, Silani V, Zecchinelli A, Perbellini L, Javoy-Agid F, et al.<br />

Clinical and pathological features in hydrocarbon-<strong>induced</strong> Parkinsonism.<br />

Ann Neurol 1996;40:922–5.<br />

Piccini P, Burn DJ, Ceravolo R, Maraganore D, Brooks DJ. The role <strong>of</strong><br />

inheritance in sporadic Parkinsons disease: evidence from a longitudinal<br />

study <strong>of</strong> dopaminergic function in twins. Ann Neurol 1999;45:577–82.<br />

Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A, Dutra A, et al.<br />

Mutation in <strong>the</strong> alpha-synuclein gene identified in families with Parkinson’s<br />

disease. Science 1997;276:2045–7.<br />

Rajput AH. Frequency and cause <strong>of</strong> Parkinson’s disease. Can J Neurol Sci<br />

1992;19:103–7.<br />

Schapira AH, Cooper JM, Dexter D, Clark JB, Jenner P, Marsden CD.<br />

Mitochondrial complex I deficiency in Parkinson’s disease. J Neurochem<br />

1990;54:823–7.<br />

Seidler A, Hellenbrand W, Robra BP, Vieregge P, Nischan P, Joerg J, et al.<br />

Possible environmental, occupational, and o<strong>the</strong>r etiologic factors for Parkinsons<br />

disease: a case-control study in Germany. Neurology 1996;46:<br />

1275–84.<br />

Semchuk KM, Love EJ, Lee RG. Parkinson’s disease and exposure to agricultural<br />

work and pesticide chemicals. Neurology 1992;42:1328–35.<br />

Semchuk KM, Love EJ, Lee RG. Parkinson’s disease: a test <strong>of</strong> <strong>the</strong> multifactorial<br />

etiologic hypo<strong>the</strong>sis. Neurology 1993;43:1173–80.<br />

Shimizu K, Ohtaki K, Matsubara K, Aoyama K, Uezono T, Saito O, et al.<br />

Carrier-mediated processes in blood-brain barrier penetration and neural<br />

uptake <strong>of</strong> <strong>paraquat</strong>. Brain Res 2001;906:135–42.<br />

Shimizu K, Matsubara K, Ohtaki K, Fujimaru S, Saito O, Shiono H. Paraquat<br />

induces long-lasting dopamine overflow through <strong>the</strong> excitotoxic pathway in<br />

<strong>the</strong> striatum <strong>of</strong> freely moving rats.. Brain Res 2003a;976:243–52.<br />

Shimizu K, Matsubara K, Ohtaki K, Shiono H. Paraquat leads to dopaminergic<br />

neural vulnerability in <strong>organ</strong>otypic midbrain culture. Neurosci Res<br />

2003b;46:523–32.<br />

Shin WW, Fong WF, Pang SF, Wong PC. Limited blood-brain barrier transport<br />

<strong>of</strong> polyamines. J Neurochem 1985;44:1056–9.<br />

Sian J, Dexter DT, Lees AJ, Daniel S, Agid Y, Javoy-Agid F, et al. Alterations in<br />

glutathione levels in Parkinson’s disease and o<strong>the</strong>r neurodegenerative<br />

disorders affecting basal ganglia. Ann Neurol 1994a;36:348–55.<br />

Sian J, Dexter DT, Lees AJ, Daniel S, Jenner P, Marsden CD. Glutathionerelated<br />

enzymes in brain in Parkinson’s disease. Ann Neurol 1994b;36:356–<br />

61.<br />

Smith LL. The identification <strong>of</strong> an accumulation system for diamines and<br />

polyamines <strong>into</strong> <strong>the</strong> lung and its relevance to <strong>paraquat</strong> toxicity. Arch Toxicol<br />

Suppl 1982;5:1–14.<br />

Spillantini MG, Schmidt ML, Lee VM, Trojanowski JQ, Jakes R, Goedert M.<br />

Alpha-synuclein in Lewy bodies. Nature 1997;388:839–40.<br />

Spillantini MG, Crow<strong>the</strong>r RA, Jakes R, Hasegawa M, Goedert M. a-Synuclein<br />

in filamentous inclusions <strong>of</strong> Lewy bodies from Parkinsons disease<br />

and dementia with Lewy bodies. Proc Natl Acad Sci USA 1998;95:6469–<br />

73.<br />

Stein WD. Kinetics <strong>of</strong> <strong>the</strong> multidrug transporter (P-glycoprotein) and its<br />

reversal. Physiol Rev 1997;77:545–90.<br />

Sveinbjornsdottir S, Hicks AA, Jonsson T, Petursson H, Gugmundsson G,<br />

Frigge ML, et al. Familial aggregation <strong>of</strong> Parkinsons disease in Iceland. N<br />

Engl J Med 2000;343:1765–70.<br />

Tanner CM. The role <strong>of</strong> environmental toxins in <strong>the</strong> etiology <strong>of</strong> Parkinson’s<br />

disease. Trends Neurosci 1989;12:49–54.<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122 1121<br />

Tanner CM, Ben-Shlomo Y. Epidemiology <strong>of</strong> Parkinson’s disease. Adv Neurol<br />

1999;80:153–9.<br />

Tanner CM, Ottman R, Goldman SM, Ellenberg J, Chan P, Mayeux R, et al.<br />

Parkinson disease in twins: an etiologic study. J Am Med Assoc 1999;281:<br />

341–6.<br />

Tawara T, Fukushima T, Hojo N, Isobe A, Shiwaku K, Setogawa T, et al. Effects<br />

<strong>of</strong> <strong>paraquat</strong> on mitochondrial electron transport system and catecholamine<br />

contents in rat brain. Arch Toxicol 1996;70:585–9.<br />

Taylor EM. The impact <strong>of</strong> efflux transporters in <strong>the</strong> brain on <strong>the</strong> development <strong>of</strong><br />

drugs for CNS disorders. Clin Pharmacokinet 2002;41:81–92.<br />

Thakar JH, Hassan MN. Effects <strong>of</strong> 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine<br />

(MPTP), cyperquat (MPP+) and <strong>paraquat</strong> on isolated mitochondria<br />

from rat striatum, cortex and liver. Life Sci 1988;43:143–9.<br />

Thiruchelvam M, Brockel BJ, Richfield EK, Baggs RB, Cory-Slechta DA.<br />

Potentiated and preferential effects <strong>of</strong> combined <strong>paraquat</strong> and maneb on<br />

nigrostriatal dopamine systems: environmental risk factors for Parkinson’s<br />

disease? Brain Res 2000a;873:225–34.<br />

Thiruchelvam M, Richfield EK, Baggs RB, Tank AW, Cory-Slechta DA. The<br />

nigrostriatal dopaminergic system as a preferential target <strong>of</strong> repeated<br />

exposures to combined <strong>paraquat</strong> and maneb: implications for Parkinson’s<br />

disease. J Neurosci 2000b;20:9207–14.<br />

Thiruchelvam M, Richfield EK, Goodman BM, Baggs RB, Cory-Slechta DA.<br />

Developmental exposure to <strong>the</strong> pesticides <strong>paraquat</strong> and maneb and <strong>the</strong><br />

Parkinson’s disease phenotype. Neurotoxicology 2002;23:621–33.<br />

Thiruchelvam M, McCormack A, Richfield EK, Baggs RB, Tank AW, Di Monte<br />

DA, et al. Age-related irreversible progressive nigrostriatal dopaminergic<br />

neurotoxicity in <strong>the</strong> <strong>paraquat</strong> and maneb model <strong>of</strong> <strong>the</strong> Parkinson’s disease<br />

phenotype. Eur J Neurosci 2003;18:589–600.<br />

Thiruchelvam MJ, Powers JM, Cory-Slechta DA, Richfield EK. Risk factors for<br />

dopaminergic neuron loss in human alpha-synuclein transgenic mice. Eur J<br />

Neurosci 2004;19:845–54.<br />

Tomita M. Comparison <strong>of</strong> one-electron reduction activity against <strong>the</strong> bipyridylium<br />

herbicides, <strong>paraquat</strong> and diquat, in microsomal and mitochondrial<br />

fractions <strong>of</strong> liver, lung and kidney (in vitro). Biochem Pharmacol 1991;42:<br />

303–9.<br />

Uitti RJ, Snow BJ, Shinotoh H, Vingerhoets FJ, Hayward M, Hashimoto S, et<br />

al. Parkinsonism <strong>induced</strong> by solvent abuse. Ann Neurol 1994;35:616–9.<br />

Uversky VN. A protein-chameleon: conformational plasticity <strong>of</strong> a-synuclein, a<br />

disordered protein involved in neurodegenerative disorders. J Biomol Struct<br />

Dyn 2003;21:211–34.<br />

Uversky VN, Li J, Fink AL. Pesticides directly accelerate <strong>the</strong> rate <strong>of</strong> alphasynuclein<br />

fibril formation: a possible factor in Parkinsons disease. FEBS<br />

Lett 2001;500:105–8.<br />

Uversky VN, Li J, Bower K, Fink AL. Synergistic effects <strong>of</strong> pesticides and<br />

metals on <strong>the</strong> fibrillation <strong>of</strong> a-synuclein: implications for Parkinsons<br />

disease. Neurotoxicology 2002;23:527–36.<br />

Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, Gispert S, et<br />

al. Hereditary early-onset Parkinson’s disease caused by mutations in<br />

PINK1. Science 2004;304:1158–60.<br />

Vanacore N, Nappo A, Gentile M, Brustolin A, Palange S, Liberati A, et al.<br />

Evaluation <strong>of</strong> risk <strong>of</strong> Parkinsons disease in a cohort <strong>of</strong> licensed pesticide<br />

users. Neurol Sci 2002;23(Suppl 2):S119–20.<br />

Widdowson PS, Farnworth MJ, Simpson MG, Lock EA. Influence <strong>of</strong> age on <strong>the</strong><br />

passage <strong>of</strong> <strong>paraquat</strong> through <strong>the</strong> blood-brain barrier in rats: a distribution<br />

and pathological examination. Hum Exp Toxicol 1996;15:231–6.<br />

Wooten GF. Neurochemistry and neuropharmacology <strong>of</strong> Parkinson’s disease.<br />

In: Watts RL, Koller W, editors. Movement disorders: neurologic principles<br />

and practice. New York: McGraw-Hill; 1997. p. 153–60.<br />

Wooten GF, Currie LJ, Bovbjerg VE, Lee JK, Patrie J. Are men at greater risk<br />

for Parkinson’s disease than women? J Neurol Neurosurg Psychiatry<br />

2004;75:637–9.<br />

Wu X, Kekuda R, Huang W, Fei YJ, Leibach FH, Chen J, et al. Identity <strong>of</strong> <strong>the</strong><br />

<strong>organ</strong>ic cation transporter OCT3 as <strong>the</strong> extraneuronal monoamine transporter<br />

(uptake2) and evidence for <strong>the</strong> expression <strong>of</strong> <strong>the</strong> transporter in <strong>the</strong><br />

brain. J Biol Chem 1998;273:32776–8.<br />

Yamada K, Fukushima T. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. II. Organ<br />

specificity <strong>of</strong> <strong>paraquat</strong>-stimulated lipid peroxidation in <strong>the</strong> inner membrane<br />

<strong>of</strong> mitochondria. Exp Toxicol Pathol 1993;45:375–80.


1122<br />

Yang W, Sun AY. Paraquat-<strong>induced</strong> free radical reaction in mouse brain<br />

microsomes. Neurochem Res 1998;23:47–53.<br />

Yasui M, Kihira T, Ota K. Calcium, magnesium and aluminum concentrations<br />

in Parkinsons disease. Neurotoxicology 1992;13:593–600.<br />

Youngman RJ, Elstner EF. Oxygen species in <strong>paraquat</strong> toxicity: <strong>the</strong> crypto-OH<br />

radical. FEBS Lett 1981;129:265–8.<br />

R.J. Dinis-Oliveira et al. / NeuroToxicology 27 (2006) 1110–1122<br />

Zarranz JJ, Alegre J, Gomez-Esteban JC, Lezcano E, Ros R, Ampuero I, et al.<br />

The new mutation, E46K, <strong>of</strong> alpha-synuclein causes Parkinson and Lewy<br />

body dementia. Ann Neurol 2004;55:164–73.<br />

Zhang L, Dresser MJ, Gray AT, Yost SC, Terashita S, Giacomini KM. Cloning<br />

and functional expression <strong>of</strong> a human liver <strong>organ</strong>ic cation transporter. Mol<br />

Pharmacol 1997;51:913–21.


_______________________________Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation<br />

PART I<br />

2. GENERAL AND SPECIFIC OBJECTIVES OF THE DISSERTATION<br />

2. GENERAL AND SPECIFIC OBJECTIVES<br />

OF THE DISSERTATION<br />

115


Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation_______________________________<br />

116


_______________________________Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation<br />

2. GENERAL AND SPECIFIC OBJECTIVES OF THE DISSERTATION<br />

Bearing on mind <strong>the</strong> above-mentioned considerations regarding to <strong>the</strong> inexistence<br />

<strong>of</strong> an antidote or effective treatment to decrease <strong>the</strong> PQ accumulation in <strong>the</strong> lung or to<br />

disrupt its toxicity, <strong>the</strong> global aim <strong>of</strong> this dissertation was to study <strong>the</strong> <strong>mechanisms</strong> <strong>of</strong><br />

PQ <strong>induced</strong> lung-toxicity and to develop and apply efficient antidotes to be used in<br />

human PQ poisonings. It is expected that an enhancement <strong>of</strong> <strong>the</strong> knowledge in this field,<br />

resulting from this dissertation, will provide new tools for medical doctors in <strong>the</strong><br />

difficult task <strong>of</strong> treating PQ <strong>into</strong>xicated patients and thus to reduce <strong>the</strong> morbidity and<br />

mortality associated to this herbicide.<br />

Some hypo<strong>the</strong>sis that derived from <strong>the</strong>se general objectives supported <strong>the</strong> specific<br />

goals <strong>of</strong> <strong>the</strong> original <strong>research</strong> corresponding to <strong>the</strong> six chapters <strong>of</strong> this <strong>the</strong>sis, as follows:<br />

CHAPTER I<br />

To study:<br />

(i) <strong>the</strong> usefulness <strong>of</strong> <strong>the</strong> isolated rat lung model when applied to characterize <strong>the</strong><br />

toxicokinetic behaviour <strong>of</strong> PQ in this tissue after bolus injection under<br />

standard experimental conditions;<br />

(ii) <strong>the</strong> influence <strong>of</strong> iso-osmotic perfusion medium replacement <strong>of</strong> Na + by Li + in<br />

<strong>the</strong> toxicokinetic parameters in <strong>the</strong> pulmonary tissue;<br />

CHAPTER II<br />

To describe:<br />

(i) a successful clinical case, regarding <strong>the</strong> <strong>into</strong>xication <strong>of</strong> a 15-year-old girl by<br />

a presumed lethal dose <strong>of</strong> PQ;<br />

117


Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation_______________________________<br />

118<br />

(ii) <strong>the</strong> status-<strong>of</strong>-<strong>the</strong>-art concerning <strong>the</strong> biochemical and toxicological aspects <strong>of</strong><br />

CHAPTER III<br />

To study:<br />

PQ poisoning and <strong>the</strong> pharmacological basis <strong>of</strong> <strong>the</strong> respective treatment.<br />

(i) , for <strong>the</strong> first time, <strong>the</strong> process <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-glycoprotein (P-gp),<br />

by DEX, in <strong>the</strong> concentration <strong>of</strong> PQ in rat lung and on its urinary and faecal<br />

excretion. The preventive effect <strong>of</strong> this pharmacological approach against<br />

PQ-<strong>induced</strong> lung toxicity was also evaluated using both biochemical and<br />

histopathological biomarkers <strong>of</strong> toxicity. Verapamil (VER), as competitive<br />

inhibitor <strong>of</strong> P-gp, was used to confirm <strong>the</strong> importance <strong>of</strong> this transporter in<br />

PQ excretion.<br />

CHAPTER IV<br />

To study:<br />

(i) <strong>the</strong> effect <strong>of</strong> DEX administration on inflammatory reaction, oxidative stress<br />

and related damage, assessed by histological and biochemical parameters in<br />

lung, liver, kidney and spleen <strong>of</strong> acute PQ-<strong>into</strong>xicated rats;<br />

(ii) <strong>the</strong> overall healing provided by DEX as well as <strong>the</strong> hypo<strong>the</strong>tic contribution<br />

<strong>of</strong> P-gp de novo syn<strong>the</strong>sis to that protection by presenting <strong>the</strong> survival rate<br />

curves.


_______________________________Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation<br />

CHAPTER V<br />

To study:<br />

(i) <strong>the</strong> role <strong>of</strong> <strong>the</strong> oxidative stress, platelet aggregation, nuclear factor (NF)-κB<br />

activation and fibrosis in PQ-<strong>induced</strong> lung toxicity;<br />

(ii) <strong>the</strong> healing effects obtained by <strong>the</strong> administration <strong>of</strong> sodium salicylate<br />

(NaSAL).<br />

CHAPTER VI<br />

To study:<br />

(i) <strong>the</strong> ability <strong>of</strong> PQ to induce apoptotic events in <strong>the</strong> lungs <strong>of</strong> Wistar rats to<br />

better understanding <strong>of</strong> <strong>the</strong> underlying adverse <strong>mechanisms</strong> <strong>induced</strong> by PQ<br />

in <strong>the</strong> respiratory tract. Firstly, <strong>the</strong> two main caspase cascades were studied<br />

through <strong>the</strong> measurement <strong>of</strong> <strong>the</strong> cytosolic cytochrome c (Cyt c)<br />

concentrations and <strong>of</strong> <strong>the</strong> enzymatic activities <strong>of</strong> caspases-1, -8 and -3, and<br />

finally, <strong>the</strong> expressions <strong>of</strong> p53 and activator protein-1 (AP-1) were<br />

evaluated, along with DNA fragmentation as a final criterion for apoptosis;<br />

(ii) <strong>the</strong> hypo<strong>the</strong>tical beneficial effects <strong>of</strong> NaSAL at this level. Overall, this<br />

should lead to a better understanding <strong>of</strong> <strong>the</strong> underlying adverse <strong>mechanisms</strong><br />

<strong>induced</strong> by PQ in <strong>the</strong> respiratory tract.<br />

119


Part I – General and Specific Objectives <strong>of</strong> <strong>the</strong> Dissertation_______________________________<br />

120


PART II<br />

____________________________________________________Part II – Original <strong>research</strong><br />

1. ORIGINAL RESEARCH<br />

121


Part I – Original <strong>research</strong>____________________________________________________<br />

122


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER I<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung:<br />

Influence <strong>of</strong> sodium depletion<br />

Reprinted from Xenobiotica 36: 724-737<br />

Copyright© (2006) with kind permission from Informa UK Ltd<br />

123


Part I – Original <strong>research</strong>____________________________________________________<br />

124


Xenobiotica, August 2006; 36(8): 724–737<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung: Influence<br />

<strong>of</strong> sodium depletion<br />

R. J. DINIS-OLIVEIRA 1,2 , M. J. DE JESÚS VALLE 2 , M. L. BASTOS 1 ,<br />

F. CARVALHO 1 , & A. SÁNCHEZ NAVARRO 2<br />

1 Faculty <strong>of</strong> Pharmacy, Department <strong>of</strong> Toxicology, University <strong>of</strong> Porto, REQUIMTE, Porto, Portugal<br />

and 2 Department <strong>of</strong> Pharmacy and Pharmaceutical Technology, Faculty <strong>of</strong> Pharmacy, University <strong>of</strong><br />

Salamanca, Salamanca, Spain<br />

(Received 20 December 2005)<br />

Abstract<br />

Paraquat accumulates in <strong>the</strong> lung through a characteristic polyamine uptake system. It has been<br />

previously shown that <strong>paraquat</strong> uptake can be significantly prevented if extracellular sodium (Na þ )is<br />

reduced, although <strong>the</strong> available data correspond to experiments performed using tissue slices or<br />

incubated cells. This type <strong>of</strong> in vitro study fails to give information on <strong>the</strong> actual behaviour occurring<br />

in vivo since <strong>the</strong> anatomy and physiology <strong>of</strong> <strong>the</strong> studied tissue is disrupted. Accordingly, <strong>the</strong> aim <strong>of</strong> <strong>the</strong><br />

present study was to explore <strong>the</strong> usefulness <strong>of</strong> <strong>the</strong> isolated rat lung model when applied to characterize<br />

<strong>the</strong> kinetic behaviour <strong>of</strong> <strong>paraquat</strong> in this tissue after bolus injection under standard experimental<br />

conditions as well as to evaluate <strong>the</strong> influence <strong>of</strong> iso-osmotic replacement <strong>of</strong> Na þ by lithium (Li þ )in<br />

<strong>the</strong> perfusion medium. The obtained results show that <strong>the</strong> present isolated rat lung model is useful for<br />

<strong>the</strong> analysis <strong>of</strong> <strong>paraquat</strong> toxicokinetics, which is reported herein for <strong>the</strong> first time. It was also observed<br />

that Na þ depletion in <strong>the</strong> perfusion medium leads to a decreased uptake <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat<br />

lung, although it seems that this condition does not contribute to improve <strong>the</strong> elimination <strong>of</strong> <strong>paraquat</strong><br />

once <strong>the</strong> herbicide reaches <strong>the</strong> extravascular structures <strong>of</strong> <strong>the</strong> tissue, since <strong>the</strong> <strong>paraquat</strong> tissue wash-out<br />

phase is similar under both experimental conditions assayed.<br />

Keywords: Isolated rat lung, <strong>paraquat</strong>, toxicokinetics, sodium, lithium<br />

Introduction<br />

Since its introduction in agriculture in 1962 (Onyeama and Oehme 1984), <strong>the</strong> widespread,<br />

non-selective contact herbicide <strong>paraquat</strong>, used as a desiccant and defoliant in a variety<br />

Correspondence: A. Sánchez Navarro, Faculty <strong>of</strong> Pharmacy, Department <strong>of</strong> Pharmacy and Pharmaceutical Technology, University<br />

<strong>of</strong> Salamanca, Avda. Campo Charro s/n. 37007, Salamanca, Spain. Tel: 34-923-294536. Fax: 34-923-294515. E-mail: asn@usal.es<br />

ISSN 0049-8254 print/ISSN 1366-5928 online ß 2006 Informa UK Ltd.<br />

DOI: 10.1080/00498250600790331


<strong>of</strong> crops, has caused thousands <strong>of</strong> deaths from both accidental and voluntary ingestion, as<br />

well as from dermal exposure. It may be considered as one <strong>of</strong> <strong>the</strong> most toxic poisons<br />

frequently used for suicide attempts. Never<strong>the</strong>less, it is readily available without restriction<br />

in several countries where it is registered. Depending on <strong>the</strong> ingested dose, different clinical<br />

patterns and outcomes have been observed in animals and humans. A large oral dose <strong>of</strong><br />

<strong>paraquat</strong> (>30 mg kg 1 in humans) rapidly leads to death from multi-<strong>organ</strong> failure, with lung<br />

damage consisting <strong>of</strong> disruption <strong>of</strong> <strong>the</strong> alveolar epi<strong>the</strong>lial cells, haemorrhage, oedema, and<br />

infiltration <strong>of</strong> inflammatory cells <strong>into</strong> <strong>the</strong> interstitial and alveolar spaces. Smaller doses <strong>of</strong><br />

<strong>paraquat</strong> (16 mg kg 1 ) may also lead to death, but this occurs after several days as a result <strong>of</strong><br />

a progressive lung fibrosis, by proliferation <strong>of</strong> fibroblasts and excessive collagen deposition<br />

showing that <strong>the</strong> main target <strong>organ</strong> for <strong>paraquat</strong> toxicity is <strong>the</strong> lung (Onyeama and Oehme<br />

1984).<br />

The direct cellular toxicity <strong>of</strong> <strong>paraquat</strong> is essentially due to its redox cycling (Figure 1):<br />

<strong>paraquat</strong> is reduced enzymatically, mainly by -nicotinamide adenine dinucleotide<br />

phosphate (NADPH)-cytochrome P450 reductase (Clejan and Cederbaum 1989) and<br />

+<br />

H3N CH2 CH2 CH2 CH2 +<br />

NH3 0.622 nm<br />

NAD(P) +<br />

NAD(P)H<br />

Putrescine<br />

H<br />

+<br />

3C N<br />

A<br />

H +<br />

3C N<br />

Redox cycle<br />

.+<br />

PQ<br />

PQ 2+<br />

.<br />

N<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 725<br />

Interstitial<br />

space<br />

CH 3<br />

+<br />

N CH3 NO .<br />

NO .<br />

O 2<br />

.-<br />

O2 ONOO -<br />

H 3 C<br />

N +<br />

SOD<br />

HWR<br />

.<br />

HO<br />

TISSUE DAMAGE<br />

0.702 nm<br />

CAT<br />

Paraquat<br />

H 2 O 2<br />

FR<br />

GSH<br />

NADPH<br />

GPX<br />

Gred<br />

N +<br />

GSSG<br />

H 2 O<br />

NADP +<br />

CH 3<br />

Cytoplasm<br />

Figure 1. Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. (A) Cellular diaphorases;<br />

SOD, superoxide dismutase or spontaneously; CAT, catalase; GPX, glutathione peroxidase; Gred,<br />

glutathione reductase; PQ 2þ , <strong>paraquat</strong>; PQ þ , <strong>paraquat</strong> cation radical; FR, Fenton reaction; HWR,<br />

Haber–Weiss reaction.


726 R. J. Dinis-Oliveira et al.<br />

NADH:ubiquinone oxidoreductase (complex I) (Fukushima et al. 1993; Yamada and<br />

Fukushima 1993) to form a <strong>paraquat</strong> monocation-free radical. The <strong>paraquat</strong> monocationfree<br />

radical is <strong>the</strong>n rapidly reoxidized in <strong>the</strong> presence <strong>of</strong> oxygen, thus resulting in <strong>the</strong><br />

generation <strong>of</strong> <strong>the</strong> superoxide radical (O2 . ) (Bus et al. 1974; Dicker and Cederbaum 1991).<br />

This <strong>the</strong>n sets in <strong>the</strong> well-known cascade leading to generation <strong>of</strong> <strong>the</strong> hydroxyl radical and<br />

consequent deleterious effects. Indeed, hydroxyl radicals (Bus et al. 1975; Youngman and<br />

Elstner 1981) have been implicated in <strong>the</strong> initiation <strong>of</strong> membrane damage by lipid<br />

peroxidation during exposure to <strong>paraquat</strong> in vitro (Bus et al. 1975) as well as in vivo<br />

(Burk et al. 1980; Dicker and Cederbaum 1991) by attack on polyunsaturated lipids, <strong>the</strong><br />

depolymerization <strong>of</strong> hyaluronic acid, <strong>the</strong> inactivation <strong>of</strong> proteins, and damage <strong>of</strong><br />

DNA. Besides, <strong>research</strong>ers have recently suggested <strong>the</strong> hypo<strong>the</strong>sis <strong>of</strong> cytotoxicity via<br />

mitochondrial dysfunction caused by <strong>paraquat</strong> (Blaszczynski et al. 1985; Hirai et al. 1985;<br />

Thakar and Hassan 1988; Tomita 1991; Fukushima et al. 1994; Tawara et al. 1996). The<br />

O2 . resulting from <strong>the</strong> <strong>paraquat</strong> redox-cycle may also react with nitric oxide (NO .)<br />

produced by nitric oxide synthase (NOS) leading to <strong>the</strong> formation <strong>of</strong> <strong>the</strong> toxic reactive<br />

species peroxynitrite anion (ONOO ) (LaVoie and Hastings 1999), thus fur<strong>the</strong>r<br />

contributing to <strong>paraquat</strong> damage.<br />

Rose et al. (1974) demonstrated that <strong>the</strong> accumulation <strong>of</strong> radioactively labelled <strong>paraquat</strong><br />

in rat lung slices was energy-dependent and obeyed saturation kinetics. O<strong>the</strong>r studies led to<br />

<strong>the</strong> conclusion that <strong>paraquat</strong> accumulated in <strong>the</strong> lung through a system for which <strong>the</strong><br />

polyamines are <strong>the</strong> natural substrates and that, in comparison with o<strong>the</strong>r <strong>organ</strong>s, <strong>the</strong> lungs,<br />

and more specifically <strong>the</strong> alveolar epi<strong>the</strong>lial cells, are endowed with a particularly active<br />

uptake system (Smith 1982; Rannels et al. 1985, 1989; Nemery et al. 1987; Dinsdale et al.<br />

1991). An important aim <strong>of</strong> earlier studies concerning pulmonary polyamine uptake was to<br />

discover <strong>the</strong> structural requirements for substrates <strong>of</strong> <strong>the</strong> transport system in order to find<br />

possible antagonists capable <strong>of</strong> preventing <strong>paraquat</strong> from entering its target cells. Ross and<br />

Krieger (1981) established that to act as a substrate for <strong>the</strong> pulmonary polyamine uptake<br />

system, a molecule must possess <strong>the</strong> following characteristics: (1) two or more positively<br />

charged nitrogen atoms, (2) maximum positivity <strong>of</strong> charge surrounding <strong>the</strong>se nitrogens,<br />

(3) a non-polar group between <strong>the</strong>se charges, and (4) a minimum <strong>of</strong> steric hindrance.<br />

Gordonsmith et al. (1983) have demonstrated that <strong>the</strong> optimum distance between <strong>the</strong><br />

nitrogen centres is four methylene groups (6.6 A ˚ ), although a spacing between four and<br />

seven methylene groups is tolerated. These data explain how polyamines and <strong>paraquat</strong> (with<br />

7.0 A ˚ between two positively charged nitrogens) can share a common uptake system, but<br />

also why <strong>paraquat</strong> (with its steric hindrance <strong>of</strong> <strong>the</strong> nitrogens by <strong>the</strong> pyridine rings) is a less<br />

successful substrate. Although <strong>paraquat</strong> proved to be a ra<strong>the</strong>r ‘poor’ substrate for <strong>the</strong><br />

polyamine uptake system, it is undoubtedly accumulated <strong>into</strong> <strong>the</strong> lung through this transport<br />

pathway.<br />

In a study performed in mice neuroblastoma cells (Rinehart and Chen 1984), <strong>the</strong><br />

possibility <strong>of</strong> sodium (Na þ ) requirement for putrescine uptake was examined by iso-osmotic<br />

replacement <strong>of</strong> Na þ by choline or lithium (Li þ ) in <strong>the</strong> incubation medium. The putrescine<br />

uptake decreased with decreasing extracellular Na þ concentration, suggesting a strong<br />

Na þ dependency <strong>of</strong> <strong>the</strong> polyamine uptake system. Na þ replacement by Li þ was also<br />

demonstrated to inhibit significantly putrescine uptake by rat isolated enterocytes (Kumagai<br />

and Johnson 1988). In bovine arterial smooth muscle cells, Janne et al. (1978) and Aziz et al.<br />

(1994) demonstrated some Na þ dependence for polyamine uptake. Rannels et al. (1989)


found that in type II pneumocytes <strong>the</strong> uptake <strong>of</strong> putrescine and spermidine was dependent<br />

on Na þ , whereas spermine uptake was not dependent on extracellular Na þ , indicating that<br />

polyamine uptake may take place via different transporters systems. Like putrescine, uptake<br />

<strong>of</strong> <strong>the</strong> herbicide <strong>paraquat</strong> was extensively inhibited as extracellular Na þ was reduced.<br />

The various studies mentioned above suggest that <strong>paraquat</strong>, like some polyamines,<br />

shows cell uptake depending on Na þ levels, particularly in type II pneumocytes.<br />

Never<strong>the</strong>less, <strong>the</strong> available data correspond to experiments performed using tissue slices<br />

or incubated cells. This latter type <strong>of</strong> in vitro studies provide information about <strong>the</strong><br />

intrinsic affinity <strong>of</strong> a compound for <strong>the</strong> incubated structure but fails to give information on<br />

<strong>the</strong> actual behaviour occurring in vivo since <strong>the</strong> anatomy and physiology <strong>of</strong> <strong>the</strong> studied<br />

tissue is disrupted.<br />

The isolated tissue and artificial perfusion techniques facilitate <strong>the</strong> performance <strong>of</strong> studies<br />

aimed at characterizing <strong>the</strong> intrinsic behaviour <strong>of</strong> a compound in a particular tissue with no<br />

interference <strong>of</strong> <strong>the</strong> rest <strong>of</strong> <strong>the</strong> body while maintaining its anatomical integrity and<br />

physiological properties. Accordingly, <strong>the</strong> aim <strong>of</strong> <strong>the</strong> present study was to explore <strong>the</strong><br />

usefulness <strong>of</strong> <strong>the</strong> isolated rat lung model when applied to <strong>the</strong> characterization <strong>of</strong> <strong>the</strong> kinetic<br />

behaviour <strong>of</strong> <strong>paraquat</strong> in this tissue under standard experimental conditions as well as to<br />

evaluate <strong>the</strong> influence <strong>of</strong> iso-osmotic replacement <strong>of</strong> Na þ by Li þ in <strong>the</strong> perfusion medium on<br />

<strong>the</strong> kinetics <strong>of</strong> this xenobiotic in pulmonary tissue.<br />

Materials and methods<br />

Reagents<br />

Paraquat dichloride (purchased as methylviologen, 98% chemical purity) and fraction V<br />

bovine serum albumin were obtained from Sigma-Aldrich (USA). O<strong>the</strong>r reagents such<br />

as NaCl, KCl, CaCl2, KH2PO4, MgSO4 7H2O, NaHCO3, NaOH, glucose, sodium<br />

dithionite, sulfosalicylic acid and lithium chloride were obtained from Panreac (Spain).<br />

Heparin 5000 UI ml –1 and sodium thiopental were obtained from B. Braun (Spain).<br />

Perfusion medium composition<br />

The perfusion medium was a modified Krebs–Henseleit bicarbonate buffer, pH 7.4,<br />

equilibrated with a carbogen mixture (95% O2, 5% CO2) and containing NaCl<br />

(119 mM), KCl (4.7 mM), CaCl2 (3.6 mM), KH2PO4 (1.18 mM), MgSO4 7H2O<br />

(1.18 mM), NaHCO3 (25 mM), glucose (5 mM) and fraction V bovine serum albumin<br />

(3%; wt/vol.).<br />

Sodium replacement<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 727<br />

The NaCl (119 mM) present in <strong>the</strong> Krebs–Henseleit medium was replaced by an equal<br />

molar concentration <strong>of</strong> LiCl.


728 R. J. Dinis-Oliveira et al.<br />

Animals<br />

The study was performed using 20 adult male rats (SLC: Wistar strain) with a mean body<br />

weight <strong>of</strong> 263.2 13.0 g. The animals were divided <strong>into</strong> two experimental groups. One<br />

group (n ¼ 10) was assigned as control group whereas ano<strong>the</strong>r (n ¼ 10) was assigned as <strong>the</strong><br />

Li þ perfusion group. The animals were maintained with water and food (standard laboratory<br />

chow) (Agway RMH-3000 chow) ad libitum until <strong>the</strong> moment <strong>of</strong> anaes<strong>the</strong>sia, which was<br />

<strong>induced</strong> with sodium thiopental (60 mg kg 1 , intraperitoneally). Housing (in Plexi-glass<br />

cages) and <strong>the</strong> experimental treatment <strong>of</strong> animals were in accordance with National<br />

Institutes <strong>of</strong> Health guidelines. The experiments complied with <strong>the</strong> current laws <strong>of</strong> Spain<br />

and Portugal.<br />

Isolated lung model: Surgical procedure<br />

The method used to isolate lungs and to keep <strong>the</strong>m artificially perfused and mechanically<br />

ventilated has been described previously (Martinez et al. 2005). Briefly, it consists <strong>of</strong> <strong>the</strong><br />

following steps:<br />

. Tracheotomy and tracheal cannulation <strong>of</strong> <strong>the</strong> animals placed in <strong>the</strong> decubito supino<br />

position on an electric blanket heated at 37 C, followed by <strong>the</strong> immediate connection <strong>of</strong><br />

<strong>the</strong> cannula to a mechanical ventilation system and subsequent injection <strong>of</strong> sodium<br />

heparin through <strong>the</strong> intraperitoneal route. The ventilation system works under positive<br />

pressure and provides warmed and moistened air to <strong>the</strong> lungs.<br />

. Opening <strong>of</strong> <strong>the</strong> thorax by two lateral transversal and one central longitudinal incision<br />

to expose <strong>the</strong> thoracic viscera.<br />

. Localization <strong>of</strong> <strong>the</strong> pulmonary and aortic arteries to place a loose ligature around both<br />

vessels in a position very close to <strong>the</strong> heart.<br />

. Incision <strong>of</strong> <strong>the</strong> left ventricle, insertion <strong>of</strong> a previously heparinized outflow cannula and<br />

fixation <strong>of</strong> <strong>the</strong> cannula by clamping.<br />

. Incision <strong>of</strong> <strong>the</strong> right ventricle, insertion <strong>of</strong> a heparinized inflow cannula and tying <strong>the</strong><br />

ligature previously placed to fix <strong>the</strong> cannula just before <strong>the</strong> bifurcation <strong>of</strong> <strong>the</strong> pulmonary<br />

artery. The inflow cannula was connected to <strong>the</strong> mechanical pump before its insertion to<br />

initiate artificial perfusion <strong>of</strong> <strong>the</strong> lungs at <strong>the</strong> same moment as <strong>the</strong> blood supply to <strong>the</strong><br />

tissue was interrupted. After a stabilization period <strong>of</strong> 5 min to wash out residual blood<br />

elements from <strong>the</strong> pulmonary circulation, <strong>paraquat</strong> was injected and a sample collection<br />

<strong>of</strong> efferent fluid was started.<br />

Experimental monitoring<br />

. Visualization <strong>of</strong> <strong>the</strong> preparation at <strong>the</strong> start <strong>of</strong> <strong>the</strong> perfusion to check that <strong>the</strong> whole lung<br />

was properly perfused. The presence <strong>of</strong> local areas with a slow washout <strong>of</strong> blood indicated<br />

deficiencies in <strong>the</strong> procedure and was a criterion for <strong>the</strong> non-viability <strong>of</strong> <strong>the</strong> experimental<br />

model.<br />

. Visualization <strong>of</strong> <strong>the</strong> preparation throughout <strong>the</strong> experimental procedure in order to verify<br />

that no tissue oedema was present. The development <strong>of</strong> translucent areas as <strong>the</strong>


experiments progressed indicated deficiencies in <strong>the</strong> procedure and was a criterion for <strong>the</strong><br />

non-viability <strong>of</strong> <strong>the</strong> experimental model.<br />

. Continuous measurement and recording <strong>of</strong> <strong>the</strong> flow rate and hydrostatic pressure at<br />

arterial level, using a probe and pressure transducer connected to <strong>the</strong> inflow cannula and<br />

to <strong>the</strong> corresponding data acquisition s<strong>of</strong>tware. A flow rate and a hydrostatic pressure<br />

out <strong>of</strong> <strong>the</strong> interval 5 0.5 ml and 13 2 mm Hg were considered as criteria <strong>of</strong> nonviability<br />

<strong>of</strong> <strong>the</strong> experimental model.<br />

Experimental equipment<br />

The main elements <strong>of</strong> <strong>the</strong> system and <strong>the</strong> experimental conditions selected were <strong>the</strong><br />

following:<br />

. Rodent ventilator (7025 Ugo Basile): this element was pre-set to supply a tidal volume <strong>of</strong><br />

2 ml at a respiratory frequency <strong>of</strong> 60 rpm with room air previously conditioned at 37 C<br />

and saturated humidity. Conditioning <strong>of</strong> <strong>the</strong> air prior to its supply to <strong>the</strong> animals was<br />

carried out by connecting <strong>the</strong> ventilator to a double-jacketed chamber <strong>into</strong> which<br />

atmospheric air was bubbled through water at 37 C. The ventilator took <strong>the</strong> air from this<br />

chamber instead <strong>of</strong> supplying non-conditioned atmospheric air.<br />

. Perfusion pump (Minipuls Õ 3 Gilson): provided a non-pulsatile flow rate <strong>of</strong> 5 ml min –1 <strong>of</strong><br />

<strong>the</strong> perfusion medium Krebs–Henseleit bicarbonate (pH 7.4) with glucose (0.9 g l –1 ) and<br />

bovine albumin (fraction V, 30 g l –1 ) in <strong>the</strong> standard and Na þ -replaced medium<br />

conditions.<br />

. Oxygenating bubbler: permits <strong>the</strong> perfusion media to be gassed effectively, with a mixture<br />

<strong>of</strong> 95% <strong>of</strong> O 2 and 5% <strong>of</strong> CO 2, 10 min prior to starting <strong>the</strong> perfusion and throughout <strong>the</strong><br />

experiments.<br />

. Bubble trap: prevents <strong>the</strong> presence <strong>of</strong> air bubbles in <strong>the</strong> medium supplied to <strong>the</strong> isolated<br />

lung.<br />

. Thermostatted bath: maintains water at 37 C circulating through <strong>the</strong> double-jacketed<br />

elements.<br />

. Fraction collector (Gilson FC 203B Fraction Collector): this was connected to <strong>the</strong><br />

outflow cannula and programmed to collect efferent fluid at <strong>the</strong> following sampling times<br />

after <strong>the</strong> dose injection: 3-s intervals for <strong>the</strong> first 1 min, and 6-s intervals over <strong>the</strong> next<br />

1 min. Subsequently, sampling time intervals were 10 s for <strong>the</strong> next 2 min; 20 s for <strong>the</strong> next<br />

2 min; 30 s for <strong>the</strong> next 2 min and 60 s for <strong>the</strong> next 12 min (a total sampling time 20 min<br />

and total samples ¼ 64).<br />

. Flow and pressure-control device: a probe (Transonic Systems, Inc. T106) was<br />

connected to <strong>the</strong> inflow cannula to measure <strong>the</strong> flow rate and corresponding pressure<br />

transducer (Transpac Õ IV, Abbott Critical Care) was fitted to determine <strong>the</strong> hydrostatic<br />

pressure at <strong>the</strong> arterial level.<br />

. Data acquisition s<strong>of</strong>tware: <strong>the</strong> Windaq (DATAQ Instruments WINDAQ, Version 1.91)<br />

program was used to record and file all <strong>the</strong> data concerning flow rate and pressure<br />

throughout <strong>the</strong> experiments.<br />

Drug injection<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 729<br />

Five minutes after <strong>the</strong> start <strong>of</strong> <strong>the</strong> artificial perfusion (stabilization period), 1000 mg <strong>of</strong><br />

<strong>paraquat</strong> dissolved in 250 ml <strong>of</strong> perfusion medium were introduced through <strong>the</strong> Y-device <strong>of</strong><br />

<strong>the</strong> inflow cannula as a bolus injection.


730 R. J. Dinis-Oliveira et al.<br />

All <strong>the</strong> experimental conditions were <strong>the</strong> same in both groups, except for Na þ levels in <strong>the</strong><br />

perfusion medium, which was replaced in one <strong>of</strong> <strong>the</strong> groups by Li þ at an iso-osmotic level.<br />

Sample pretreatment<br />

After 20 min <strong>of</strong> perfusion and sampling collection, all major cartilaginous airways were<br />

dissected free and <strong>the</strong> wet weight <strong>of</strong> <strong>the</strong> remaining lung tissue was determined. The lungs<br />

were <strong>the</strong>n processed for <strong>the</strong> measurement <strong>of</strong> <strong>the</strong> remaining <strong>paraquat</strong>. The lung tissue was<br />

homogenized (Pro 250 Homogenizer) in 50 mM phosphate buffer/0.1% Triton X-100 (pH<br />

7.4). The homogenate was kept on ice and <strong>the</strong>n centrifuged at 13 000g, 4C for 20 min.<br />

Aliquots <strong>of</strong> <strong>the</strong> resulting supernatants were treated with sulfosalicylic acid (5% in final<br />

volume) and centrifuged (13 000g, 4 C for 5 min). Aliquots <strong>of</strong> <strong>the</strong> outflow samples were also<br />

treated with sulfosalicylic acid (5% in final volume) to precipitate <strong>the</strong> proteins and<br />

centrifuged (13 000g, 4 C for 5 min). After deproteinization, <strong>the</strong> supernatants were<br />

alkalinized with 10 N NaOH (pH > 9) and <strong>the</strong>n gently mixed with <strong>the</strong> reductant (sodium<br />

dithionite) to give a blue colour characteristic <strong>of</strong> <strong>the</strong> <strong>paraquat</strong> radical. In two animals <strong>of</strong> each<br />

group <strong>the</strong> perfusion was prolonged until 30 min in order to evaluate <strong>the</strong> effect <strong>of</strong> this<br />

substitution in <strong>the</strong> viability criteria referred to above (hydrostatic pressure increase and<br />

appearance <strong>of</strong> translucent areas).<br />

Paraquat analysis<br />

Paraquat quantitation in outflow perfusate and lung homogenates was carried out using a<br />

rapid, simple method based on second-derivative spectrophotometry (Fell et al. 1981; Fuke<br />

et al. 1992; Kuo et al. 2001) using a Shimadzu model UV/VIS 160 double-beam with a<br />

built-in microcomputer and a quartz cell with an optical path length <strong>of</strong> 1.0 cm.<br />

No interference was observed in <strong>the</strong> zero-order and second-derivative spectrum <strong>of</strong> <strong>the</strong><br />

blank. The data <strong>of</strong> a zero-order spectrum obtained by scanning from 500 to 380 nm with a<br />

0.5-nm bandwidth were stored in <strong>the</strong> machine and <strong>the</strong>n differentiated with 4 nm <strong>of</strong><br />

differential wavelength to give a second derivative spectrum. A qualitative and quantitative<br />

analysis <strong>of</strong> reduced <strong>paraquat</strong> was made at <strong>the</strong> amplitude peaks <strong>of</strong> 396–403 nm <strong>of</strong> <strong>the</strong> secondderivative<br />

spectrum. The calibration curve in <strong>the</strong> 0.2–8.0 mgml –1 range obeys Beer’s law.<br />

The samples were diluted in order to fall <strong>into</strong> <strong>the</strong> reference range <strong>of</strong> <strong>the</strong> standard curve.<br />

Using <strong>the</strong>se experimental conditions, <strong>the</strong> intra- and inter-day coefficients <strong>of</strong><br />

variation showed values lower than 5% and <strong>the</strong> detection limit <strong>of</strong> <strong>the</strong> method was<br />

100 ng ml –1 .<br />

Results for <strong>the</strong> lungs were expressed in mmol PQ mg –1 protein. Lung protein quantification<br />

was performed according to <strong>the</strong> method <strong>of</strong> Lowry et al. (1951), using bovine serum<br />

albumin as standard.<br />

Toxicokinetic analysis<br />

Paraquat concentration curves in <strong>the</strong> lung efferent fluid were analysed by <strong>the</strong> statistical<br />

moment <strong>the</strong>ory (Yamaoka et al. 1978). According to this <strong>the</strong>ory, <strong>the</strong> area under <strong>the</strong> curve<br />

(AUC), mean transit time (MTT) and variance <strong>of</strong> transit time (VTT) may be estimated<br />

in <strong>the</strong> isolated lung from <strong>the</strong> stochastic analysis <strong>of</strong> <strong>the</strong> outflow concentration curve (Ct )


using <strong>the</strong> following equations:<br />

VTT ¼<br />

AUC 1 0 ¼<br />

Z 1<br />

0<br />

MTT ¼<br />

Z 1<br />

0<br />

Z 1<br />

0Z<br />

1<br />

0<br />

CðtÞdt ð1Þ<br />

t CðtÞdt<br />

CðtÞdt<br />

ðt MTTÞ 2 CðtÞdt<br />

Z 1<br />

0<br />

CðtÞdt<br />

Since <strong>the</strong> experimental system used here included tubing besides <strong>the</strong> isolated tissue, it was<br />

necessary to correct for <strong>the</strong> influence <strong>of</strong> <strong>the</strong>se devices on <strong>the</strong> MTT estimated from <strong>the</strong> above<br />

equation. Additional experiments performed under <strong>the</strong> same experimental conditions as<br />

described above, but in absence <strong>of</strong> <strong>the</strong> tissue were carried out to quantify <strong>the</strong> mean transit<br />

time <strong>of</strong> <strong>the</strong> drug in <strong>the</strong> devices (MTTd), in order to estimate <strong>the</strong> actual mean transit time <strong>of</strong><br />

<strong>the</strong> drug in <strong>the</strong> tissue (MTTa) by applying <strong>the</strong> following correction:<br />

MTTa ¼ MTT MTTd ð4Þ<br />

Assuming <strong>the</strong> experimental preparation as an stationary system, <strong>the</strong> volume <strong>of</strong> distribution<br />

<strong>of</strong> <strong>the</strong> drug in <strong>the</strong> lung (Vd) was calculated from <strong>the</strong> mean transit time <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong><br />

tissue (MTTa) and <strong>the</strong> perfusion flow rate (Q ¼ 5 ml min –1 ), as follows (Weiss 1995):<br />

ð2Þ<br />

ð3Þ<br />

Vd ¼ MTTa Q ð5Þ<br />

The distribution coefficient (V d/L w) was also calculated, L w being <strong>the</strong> weight <strong>of</strong> <strong>the</strong> isolated<br />

lung. Finally, <strong>the</strong> washout rate constant (K w) was estimated from <strong>the</strong> slope value <strong>of</strong> <strong>the</strong><br />

terminal phase <strong>of</strong> <strong>the</strong> outflow curve.<br />

Statistical analysis<br />

Results are given as <strong>the</strong> mean standard deviation (SD). Statistical comparison <strong>of</strong><br />

parameters obtained for <strong>paraquat</strong> in both groups was performed by a Student’s t-test for<br />

non-paired samples using <strong>the</strong> STATGRAPHICS Plus 4.0 program. The level <strong>of</strong> significance<br />

was set at p < 0.05.<br />

Results<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 731<br />

Figure 2 shows <strong>the</strong> mean concentration curves <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> efferent fluid for control and<br />

Na þ -depleted medium groups with <strong>the</strong> corresponding standard deviations. Very similar<br />

concentration pr<strong>of</strong>iles in <strong>the</strong> outflow fluid are observed, although some interesting differences<br />

must be highlighted. In fact, a peak value is rapidly achieved in both groups. Never<strong>the</strong>less, <strong>the</strong><br />

mean value <strong>of</strong> <strong>the</strong> peak concentration reaches a higher value in <strong>the</strong> Li þ group (679.62 43.04<br />

vs. 491.59 29.00 mgml –1 ; p < 0.001). This result is supported and confirmed by <strong>the</strong> results


732 R. J. Dinis-Oliveira et al.<br />

Q remanider (mg)<br />

900<br />

750<br />

600<br />

450<br />

300<br />

150<br />

Concentration (mg/ml)<br />

700<br />

600<br />

Control<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Lithium<br />

0.05 0.4 0.75 1.2 1.9 3 4.33 7<br />

Time (min)<br />

Figure 2. Concentration curves <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> efferent fluid <strong>of</strong> <strong>the</strong> isolated rat lung preparation<br />

after a bolus injection at 1000 mg for <strong>the</strong> control and LiCl groups. Mean values and SD are plotted, and<br />

are derived from 48 to 55 determinations for each preparation.<br />

Control<br />

Lithium<br />

0<br />

0.05 0.55 1.10 2.17<br />

Time (min)<br />

3.83 7.50<br />

Figure 3. Paraquat concentrations remaining in <strong>the</strong> lung <strong>of</strong> control and LiCl groups after a bolus<br />

injection at 1000 mg. Mean values and SD are plotted, and are derived from 48 to 55 determinations<br />

for each preparation. ***Significant differences at p < 0.001 using a Student’s t-test for non-paired<br />

samples to compare <strong>the</strong> two groups relatively to <strong>the</strong> remaining <strong>paraquat</strong> quantity after 20 min <strong>of</strong><br />

perfusion.<br />

obtained from <strong>the</strong> quantitation <strong>of</strong> remaining <strong>paraquat</strong> concentrations in <strong>the</strong> isolated lung at<br />

<strong>the</strong> end <strong>of</strong> <strong>the</strong> experiments (0.32 0.04 and 0.18 0.02 mgg 1 for standard and Na þ -<br />

depleted media, respectively; p < 0.001). The same result was obtained plotting <strong>the</strong> quantity<br />

<strong>of</strong> <strong>paraquat</strong> remaining in <strong>the</strong> lung against time (Figure 3) and shows that Na þ depletion<br />

reduces <strong>paraquat</strong> accumulation by decreasing its access to extravascular structures <strong>of</strong> <strong>the</strong> lung<br />

tissues (p < 0.001). Moreover, after 30 min <strong>of</strong> perfusion, <strong>the</strong> Na þ substitution by Li þ also<br />

conferred a substantial protection against <strong>paraquat</strong>-<strong>induced</strong> lung oedema, as observed by <strong>the</strong><br />

development <strong>of</strong> translucent areas (Figure 4) as well as by alterations in <strong>the</strong> hydrostatic<br />

pressure at arterial level (13 2 vs. 20 7 mm Hg). The differences found in <strong>the</strong> outflow<br />

curves are reflected in <strong>the</strong> parameters estimated by <strong>the</strong> statistical analysis <strong>of</strong> <strong>the</strong>se curves.<br />

Table I includes <strong>the</strong> mean values <strong>of</strong> estimated parameters, by stochastic methods,<br />

corresponding to standard and Na þ -depleted media, respectively. Although all parameters<br />

are modified in <strong>the</strong> experiments with <strong>the</strong> iso-osmotic replacement <strong>of</strong> Na þ by Li þ , only <strong>the</strong><br />

mean values <strong>of</strong> AUC and Vd/Lw show statistically significant differences (p < 0.01). The mean<br />

***


Table I. Mean parameters estimated from outflow curves <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung as evaluated<br />

by stochastic methods.<br />

Parameter<br />

AUC0–1 (mg min ml –1 ) a<br />

MTT (min) b<br />

VTT c<br />

Kw (min –1 ) d<br />

Vd (ml) e<br />

Vd/Lw (ml g –1 ) f<br />

Cmax (mgml –1 ) g<br />

AUC for <strong>the</strong> efferent fluid increases, whereas <strong>the</strong> distribution coefficient in <strong>the</strong> lung (Vd/Lw)<br />

decreases, both changes revealing a lower exposure <strong>of</strong> <strong>the</strong> tissue to <strong>paraquat</strong> when <strong>the</strong> isolated<br />

lung was perfused under Na þ -depleted conditions. In fact, <strong>the</strong> mean value <strong>of</strong> MTTa, which is<br />

<strong>the</strong> first moment <strong>of</strong> <strong>the</strong> curve that constitutes an excellent indicator <strong>of</strong> tissue exposure (Weiss<br />

and Roberts 1996), also decreases from 0.74 0.07 to 0.54 0.10 min, although <strong>the</strong><br />

statistical comparison failed to show significant differences.<br />

Discussion<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 733<br />

Figure 4. Macroscopic lung examination <strong>of</strong> <strong>the</strong> control and LiCl groups after a bolus injection <strong>of</strong><br />

1000 mg <strong>of</strong> <strong>paraquat</strong>. Photographs were taken 30 min after <strong>paraquat</strong> injection. Arrows indicate<br />

translucent areas as a result <strong>of</strong> lung oedema development.<br />

Control group (n ¼ 8),<br />

mean SD 8<br />

LiCl group (n ¼ 8),<br />

mean SD<br />

184.25 15.85 219.51 17.51**<br />

0.74 0.07 0.54 0.1<br />

1.75 0.52 0.64 0.36<br />

0.40 0.06 0.63 0.33<br />

3.70 0.36 2.69 0.52<br />

2.99 0.38 2.21 0.4**<br />

491.59 29.05 679.62 43.04***<br />

a Area under <strong>the</strong> curve.<br />

b Mean transit time.<br />

c Variance <strong>of</strong> mean transit time.<br />

d Washout rate constant.<br />

e Apparent volume <strong>of</strong> distribution.<br />

f Distribution coefficient.<br />

g Maximum concentration reached in <strong>the</strong> outflow samples.<br />

Data are means standard deviation, derived from 48 to 55 determinations for each preparation.<br />

Significant differences at **p < 0.01 and ***p < 0.001, respectively, using a Student’s t-test for non-paired samples<br />

to compare <strong>the</strong> two groups.<br />

Despite <strong>paraquat</strong> being <strong>the</strong> most toxic herbicide marketed over <strong>the</strong> last 60 years, it is <strong>the</strong><br />

third most widely used product in <strong>the</strong> world for this purpose (Wesseling et al. 1997, 2001).


734 R. J. Dinis-Oliveira et al.<br />

Nowadays, no antidote or effective treatment for <strong>paraquat</strong> poisoning has been identified,<br />

survival depending on <strong>the</strong> amount ingested and <strong>the</strong> time elapsed until beginning intensive<br />

medical measures to inactivate and eliminate <strong>paraquat</strong>. Never<strong>the</strong>less, attempts to elucidate<br />

<strong>the</strong> tissue-uptake mechanism to interfere with <strong>the</strong> process and to avoid <strong>the</strong> final toxicological<br />

consequences are being undertaken.<br />

The main objective <strong>of</strong> <strong>the</strong> present study was to investigate <strong>the</strong> kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong><br />

isolated rat lung model. We have determined <strong>the</strong> kinetic parameters and <strong>the</strong> influence <strong>of</strong><br />

Li þ in <strong>the</strong> observed kinetics. The characterization <strong>of</strong> <strong>the</strong> kinetic pr<strong>of</strong>ile <strong>of</strong> xenobiotics in<br />

specific <strong>organ</strong>s or tissues is becoming increasingly interesting since pharmacological and<br />

toxicological responses are much more related to <strong>the</strong> concentrations at target sites than in<br />

plasma. This is even more important in <strong>the</strong> case <strong>of</strong> <strong>paraquat</strong> because <strong>the</strong> kinetics are an<br />

important issue in <strong>into</strong>xication by this compound. The techniques <strong>of</strong> tissue isolation and<br />

perfusion <strong>of</strong>fer an excellent alternative to traditional methodology when detailed information<br />

about <strong>the</strong> drug distribution in a particular body tissue is required. This technique allows <strong>the</strong><br />

characterization <strong>of</strong> <strong>the</strong> kinetic pr<strong>of</strong>ile for a tissue in a single animal and avoids <strong>the</strong> interindividual<br />

variability in each single curve, leading to a corresponding reduction in curve<br />

replicates and hence a substantial reduction in <strong>the</strong> number <strong>of</strong> animals used (5–8 vs. 50–80<br />

per tissue). The results <strong>of</strong> <strong>the</strong> comparative study on <strong>the</strong> disposition <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong><br />

presence <strong>of</strong> different media compositions in <strong>the</strong> isolated rat lung confirm <strong>the</strong>se advantages<br />

and show that <strong>the</strong> data provided by this experimental model afford very useful information<br />

about <strong>the</strong> kinetic behaviour <strong>of</strong> xenobiotics in this tissue. Indeed, by using <strong>the</strong> perfused <strong>organ</strong>,<br />

compounds <strong>of</strong> interest are delivered to <strong>the</strong> structurally intact lung via <strong>the</strong> pulmonary<br />

circulation.<br />

The <strong>paraquat</strong> outflow concentration curves (Figure 2) obtained for <strong>the</strong> control group<br />

confirms <strong>the</strong> high pulmonary affinity <strong>of</strong> this compound. The polyexponential pr<strong>of</strong>ile shown<br />

by <strong>the</strong> curves reveals rapid access to extravascular spaces with a slow washout process. After<br />

reaching <strong>the</strong> peak value, <strong>the</strong> pr<strong>of</strong>ile shows a three-phase decay, each phase presumably<br />

representing <strong>the</strong> washout <strong>of</strong> <strong>the</strong> product from <strong>the</strong> vascular, interstitial and intracellular<br />

spaces, respectively. Curves (Figure 2) corresponding to <strong>the</strong> experiments performed under<br />

Na þ -depleted and standard conditions show a similar pr<strong>of</strong>ile for both efferent fluids,<br />

although <strong>the</strong> peak concentration reaches a significantly higher value in <strong>the</strong> Na þ -depleted<br />

media.<br />

This latter observation leads to <strong>the</strong> conclusion that tissue uptake <strong>of</strong> <strong>paraquat</strong> is reduced<br />

when <strong>the</strong> iso-osmotic replacement <strong>of</strong> NaCl by LiCl is performed. Such a conclusion is also<br />

supported and confirmed by <strong>the</strong> results obtained from <strong>the</strong> quantitation <strong>of</strong> <strong>paraquat</strong><br />

remaining in <strong>the</strong> lungs <strong>of</strong> both groups at <strong>the</strong> end <strong>of</strong> <strong>the</strong> experiments and from <strong>the</strong> premature<br />

oedema development in <strong>the</strong> control compared with <strong>the</strong> perfusion in absence <strong>of</strong> Na þ<br />

(Figure 4). The early appearance <strong>of</strong> oedema leads to an increase in hydrostatic pressure at<br />

<strong>the</strong> arterial level reflecting a resistance to perfusion as a consequence <strong>of</strong> water accumulation<br />

in <strong>the</strong> lung. The slope <strong>of</strong> <strong>the</strong> different exponential phases can be estimated, <strong>the</strong> value for <strong>the</strong><br />

terminal one being much lower than those corresponding to earlier phases. This may be<br />

interpreted as <strong>the</strong> existence <strong>of</strong> an inefficient export process <strong>of</strong> <strong>paraquat</strong> from <strong>the</strong> tissue.<br />

Never<strong>the</strong>less, an efficient lung polyamine uptake system used by <strong>paraquat</strong> to be transported<br />

<strong>into</strong> <strong>the</strong> alveolar lung epi<strong>the</strong>lium seems to be operating since <strong>the</strong> product accumulates and<br />

reaches a much higher concentration in <strong>the</strong> lung than in <strong>the</strong> outflow media. Ten minutes<br />

after starting sample collection no <strong>paraquat</strong> was detected in <strong>the</strong> outflow samples in both<br />

groups. In spite <strong>of</strong> this finding, <strong>paraquat</strong> was found in lung tissue after 20 min <strong>of</strong> perfusion.<br />

Again, this may reflect <strong>the</strong> difficulty for <strong>paraquat</strong> to return to <strong>the</strong> vascular space after<br />

reaching <strong>the</strong> extravascular space (especially <strong>the</strong> alveolar epi<strong>the</strong>lium). In general, a lower


tissue exposure may be due to ei<strong>the</strong>r a more restricted access <strong>of</strong> <strong>the</strong> product to <strong>the</strong> lung or<br />

to a more rapid wash-out from <strong>the</strong> intracellular compartments. In this experiment, <strong>the</strong><br />

significant increase <strong>of</strong> <strong>the</strong> AUC value toge<strong>the</strong>r with <strong>the</strong> higher peak value (in <strong>the</strong> Na þ -<br />

depleted medium conditions) suggests a more restricted access <strong>of</strong> <strong>paraquat</strong> under Na þ -<br />

depleted conditions, a conclusion also supported by <strong>the</strong> significant decrease <strong>of</strong> <strong>the</strong><br />

distribution coefficient <strong>of</strong> <strong>paraquat</strong>. The present study resulted in a peak concentration <strong>of</strong><br />

about 500 mg l –1 (¼1.95 mmol l –1 ) in <strong>the</strong> effluent perfusate for <strong>the</strong> standard conditions and a<br />

value <strong>of</strong> 0.21 mmol l –1 for <strong>the</strong> saturable <strong>paraquat</strong> uptake transport constant (K m) has been<br />

reported in lung slices (Ross and Krieger 1981). Accordingly, <strong>the</strong> uptake transport system<br />

was probably saturated under our experimental conditions.<br />

Although statistical comparison failed to show significant differences in <strong>the</strong> MTTa<br />

between <strong>the</strong> groups, <strong>the</strong> MTTa in <strong>the</strong> lungs for <strong>the</strong> standard conditions was higher, possibly<br />

implying a longer and more intense exposure <strong>of</strong> <strong>the</strong> tissue to <strong>paraquat</strong>. These data were<br />

corroborated by <strong>the</strong> values <strong>of</strong> <strong>the</strong> distribution coefficients (Vd/Lw), which also decreased in<br />

<strong>the</strong> Li þ medium (2.21 ml g –1 ) in comparison with <strong>the</strong> standard conditions (2.99 ml g –1 ).<br />

Despite a sustained effort over <strong>the</strong> last decade, <strong>the</strong> mammalian polyamine transporter has<br />

not yet been cloned and characterized. Considering <strong>the</strong> reported mechanism <strong>of</strong><br />

accumulation <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> lung tissue, and some reports that points to a Na þ<br />

dependence <strong>of</strong> <strong>the</strong> polyamine uptake system (Janne et al. 1978; Rinehart and Chen 1984;<br />

Kumagai and Johnson 1988; Rannels et al. 1989; Aziz et al. 1994), it might be suggested that<br />

<strong>paraquat</strong> uptake <strong>into</strong> <strong>the</strong> extravascular compartments <strong>of</strong> <strong>the</strong> lung (alveolar interstitium and<br />

epi<strong>the</strong>lium) is prevented by Na þ depletion. In contrast, some reports evidence no Na þ<br />

requirement for putrescine uptake by different cell types (Lewis et al. 1989; M<strong>organ</strong> 1992).<br />

The present study <strong>of</strong>fers new data by demonstrating that <strong>the</strong> polyamine uptake system for<br />

<strong>paraquat</strong> is in fact Na þ -dependent. However, o<strong>the</strong>r polyamine uptake systems exist (Rannels<br />

et al. 1989) for which competition studies have shown that for <strong>the</strong>ir substrate uptake <strong>the</strong>re is<br />

no Na þ dependence (Smith and Wyatt 1981). It must be also considered that <strong>the</strong> lung is a<br />

heterogeneous tissue comprised <strong>of</strong> approximately 40 different cell types (Sorokin 1970).<br />

Therefore, <strong>the</strong> total tissue uptake observed in <strong>the</strong> lung reflects <strong>the</strong> mean uptake activity <strong>of</strong> all<br />

<strong>the</strong> constitutive cell types present in <strong>the</strong> tissue.<br />

Taken collectively, it can be concluded that <strong>the</strong> kinetic behaviour <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated<br />

lung seems to be modified by Na þ depletion in <strong>the</strong> perfusion medium. Accordingly, although<br />

it seems that this condition does not significantly contribute to improve <strong>the</strong> elimination <strong>of</strong><br />

<strong>paraquat</strong> from <strong>the</strong> tissue once <strong>the</strong> product gets to <strong>the</strong> deepest structures, we suggest that an<br />

impaired access to <strong>the</strong> lung tissue might be operating under Na þ -depletion conditions.<br />

Acknowledgements<br />

This work is part <strong>of</strong> a Research Project (PM 1998-0138) funded by <strong>the</strong> Spanish Council<br />

for Science and Technology. Ricardo Dinis acknowledges <strong>the</strong> FCT for his PhD grant<br />

(SFRH/BD/13707/2003).<br />

References<br />

Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 735<br />

Aziz SM, Lipke DW, Olson JW, Gillespie MN. 1994. Role <strong>of</strong> ATP and sodium in polyamine transport in bovine<br />

pulmonary artery smooth cells. Biochemistry and Pharmacology 48:1611–1618.


736 R. J. Dinis-Oliveira et al.<br />

Blaszczynski M, Litwinska J, Zaborowska D, Bilinski T. 1985. The role <strong>of</strong> respiratory chain in <strong>paraquat</strong> toxicity in<br />

yeast. Acta Microbiologica Polonica 34:243–254.<br />

Burk RF, Lawrence RA, Lane JM. 1980. Liver necrosis and lipid peroxidation in <strong>the</strong> rat as a result <strong>of</strong> <strong>paraquat</strong> and<br />

diquat administration. Effect <strong>of</strong> selenium deficiency. Journal <strong>of</strong> Clinical Investigation 65:1024–1031.<br />

Bus JS, Aust SD, Gibson JE. 1974. Superoxide- and singlet oxygen-catalyzed lipid peroxidation as a possible<br />

mechanism for <strong>paraquat</strong> (methyl viologen) toxicity. Biochemical and Biophysical Research Communications<br />

58:749–755.<br />

Bus JS, Aust SD, Gibson JE. 1975. Lipid peroxidation: A possible mechanism for <strong>paraquat</strong> toxicity. Research<br />

Communications in Chemistry, Pathology and Pharmacology 11:31–38.<br />

Clejan L, Cederbaum AI. 1989. Synergistic interaction between NADPH-cytochrome P-450 reductase, <strong>paraquat</strong><br />

and iron in <strong>the</strong> generation <strong>of</strong> active oxygen radicals. Biochemistry and Pharmacology 38:1779–1786.<br />

Dicker E, Cederbaum AI. 1991. NADH-dependent generation <strong>of</strong> reactive oxygen species by microsomes in <strong>the</strong><br />

presence <strong>of</strong> iron and redox cycling agents. Biochemistry and Pharmacology 42:529–535.<br />

Dinsdale D, Preston SG, Nemery B. 1991. Effects <strong>of</strong> injury on [3H]putrescine uptake by types I and II cells in rat<br />

lung slices. Experimental Molecular Pathology 54:218–229.<br />

Fell AF, Jarvie DR, Stewart MJ. 1981. Analysis for <strong>paraquat</strong> by second- and fourth-derivative spectroscopy. Clinical<br />

Chemistry 27:286–292.<br />

Fuke C, Ameno K, Ameno S, Kiriu T, Shinohara T, Sogo K, Ijiri I. 1992. A rapid, simultaneous determination <strong>of</strong><br />

<strong>paraquat</strong> and diquat in serum and urine using second-derivative spectroscopy. Journal <strong>of</strong> Analytical Toxicology<br />

16:214–216.<br />

Fukushima T, Yamada K, Hojo N, Isobe A, Shiwaku K, Yamane Y. 1994. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>.<br />

III. The effects <strong>of</strong> acute <strong>paraquat</strong> exposure on <strong>the</strong> electron transport system in rat mitochondria. Experiments in<br />

Toxicology and Pathology 46:437–441.<br />

Fukushima T, Yamada K, Isobe A, Shiwaku K, Yamane Y. 1993. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>.<br />

I. NADH oxidation and <strong>paraquat</strong> radical formation via complex I. Experiments in Toxicology and Pathology<br />

45:345–349.<br />

Gordonsmith RH, Brooke-Taylor S, Smith LL, Cohen GM. 1983. Structural requirements <strong>of</strong> compounds to<br />

inhibit pulmonary diamine accumulation. Biochemistry and Pharmacology 32:3701–3709.<br />

Hirai K, Witschi H, Cote MG. 1985. Mitochondrial injury <strong>of</strong> pulmonary alveolar epi<strong>the</strong>lial cells in acute <strong>paraquat</strong><br />

<strong>into</strong>xication. Experimental Molecular Pathology 43:242–252.<br />

Janne J, Poso H, Raina A. 1978. Polyamines in rapid growth and cancer. Biochimica et Biophysica Acta<br />

473:241–293.<br />

Kumagai J, Johnson LR. 1988. Characteristics <strong>of</strong> putrescine uptake in isolated rat enterocytes. American Journal <strong>of</strong><br />

Physiology 254:G81–G6.<br />

Kuo TL, Lin DL, Liu RH, Moriya F, Hashimoto Y. 2001. Spectra interference between diquat and <strong>paraquat</strong> by<br />

second derivative spectrophotometry. Forensic Science International 121:134–139.<br />

LaVoie MJ, Hastings TG. 1999. Peroxynitrite- and nitrite-<strong>induced</strong> oxidation <strong>of</strong> dopamine: Implications for nitric<br />

oxide in dopaminergic cell loss. Journal <strong>of</strong> Neurochemistry 73:2546–2554.<br />

Lewis CP, Haschek WM, Wyatt I, Cohen GM, Smith LL. 1989. The accumulation <strong>of</strong> cystamine and its<br />

metabolism to taurine in rat lung slices. Biochemical Pharmacology 38:481–488.<br />

Lowry OHN, Rosebrough NJ, Farr AL, Randall RJ. 1951. Protein measurement with Folin phenol reagent. Journal<br />

<strong>of</strong> Biology and Chemistry 193:265–275.<br />

Martinez MSM, Gandarillas CIC, Lanao JM, Sanchez Navarro A. 2005. Influence <strong>of</strong> flow rate on <strong>the</strong> disposition <strong>of</strong><br />

lev<strong>of</strong>loxacin and netilmicin in <strong>the</strong> isolated rat lung. European Journal <strong>of</strong> Pharmaceutical Science 24:325–332.<br />

M<strong>organ</strong> DM. 1992. Uptake <strong>of</strong> polyamines by human endo<strong>the</strong>lial cells. Characterization and lack <strong>of</strong> effect <strong>of</strong><br />

agonists <strong>of</strong> endo<strong>the</strong>lial function. Biochemical Journal 286:413–417.<br />

Nemery B, Smith LL, Aldridge WN. 1987. Putrescine and 5-hydroxytryptamine accumulation in rat lung slices:<br />

Cellular localization and responses to cell-specific lung injury. Toxicology and Applied Pharmacology<br />

91:107–120.<br />

Onyeama HP, Oehme FW. 1984. A literature review <strong>of</strong> <strong>paraquat</strong> toxicity. Veterinary and Human Toxicology<br />

26:494–502.<br />

Rannels DE, Kameji R, Pegg AE, Rannels SR. 1989. Spermidine uptake by type II pneumocytes: Interactions<br />

<strong>of</strong> amine uptake pathways. American Journal <strong>of</strong> Physiology, Lung Cell and Molecular Physiology<br />

257:L346–L353.<br />

Rannels DE, Pegg AE, Clark RS, Addison JL. 1985. Interaction <strong>of</strong> <strong>paraquat</strong> and amine uptake by rat lungs perfused<br />

in situ. American Journal <strong>of</strong> Physiology, Endocrinology and Metabolism 249:E506–E513.<br />

Rinehart CAJ, Chen KY. 1984. Characterization <strong>of</strong> <strong>the</strong> polyamine transport system in mouse neuroblastoma cells.<br />

Effects <strong>of</strong> sodium and system A amino acids. Journal <strong>of</strong> Biology and Chemistry 259:4750–4756.


Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung 737<br />

Rose MS, Smith LL, Wyatt I. 1974. Evidence for energy-dependent accumulation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong> rat lung. Nature<br />

252:314–315.<br />

Ross JH, Krieger RI. 1981. Structure–activity correlations <strong>of</strong> amines inhibiting active uptake <strong>of</strong> <strong>paraquat</strong> (methyl<br />

viologen) <strong>into</strong> rat lung slices. Toxicology and Applied Pharmacology 59:238–249.<br />

Smith LL. 1982. The identification <strong>of</strong> an accumulation system for diamines and polyamines <strong>into</strong> <strong>the</strong> lung and its<br />

relevance to <strong>paraquat</strong> toxicity. Archives in Toxicology Suppl 5:1–14.<br />

Smith LL, Wyatt I. 1981. The accumulation <strong>of</strong> putrescine <strong>into</strong> slices <strong>of</strong> rat lung and brain and its relationship to <strong>the</strong><br />

accumulation <strong>of</strong> <strong>paraquat</strong>. Biochemistry and Pharmacology 30:1053–1058.<br />

Sorokin SP. 1970. The cells <strong>of</strong> <strong>the</strong> lungs. In: Nettesheim P, Hanna MG, Dea<strong>the</strong>ridge JW, editors. Morphology <strong>of</strong><br />

experimental respiratory carcinogenesis. Washington, DC: US Atomic Energy Commission.<br />

Tawara T, Fukushima T, Hojo N, Isobe A, Shiwaku K, Setogawa T, Yamane Y. 1996. Effects <strong>of</strong> <strong>paraquat</strong> on<br />

mitochondrial electron transport system and catecholamine contents in rat brain. Archives in Toxicology<br />

70:585–589.<br />

Thakar JH, Hassan MN. 1988. Effects <strong>of</strong> 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), cyperquat<br />

(MPPþ) and <strong>paraquat</strong> on isolated mitochondria from rat striatum, cortex and liver. Life Sciences 43:143–149.<br />

Tomita M. 1991. Comparison <strong>of</strong> one-electron reduction activity against <strong>the</strong> bipyridylium herbicides, <strong>paraquat</strong> and<br />

diquat, in microsomal and mitochondrial fractions <strong>of</strong> liver, lung and kidney (in vitro). Biochemistry and<br />

Pharmacology 42:303–309.<br />

Weiss M. 1995. Distribution kinetics in <strong>the</strong> body and single <strong>organ</strong>s: Moment analysis. In: D’Argenio DZ, editor.<br />

Advanced methods <strong>of</strong> pharmacokinetics and pharmacodynamic system analysis. New York, NY: Plenum.<br />

pp 89–100.<br />

Weiss M, Roberts MS. 1996. Tissue distribution kinetics as determinant <strong>of</strong> transit time dispersion <strong>of</strong> drugs in<br />

<strong>organ</strong>s: Application <strong>of</strong> a stochastic model to <strong>the</strong> rat hindlimb. Journal <strong>of</strong> Pharmacokinetics and Biopharmacology<br />

24:173–196.<br />

Wesseling C, Hogstedt C, Picado A, Johansson L. 1997. Unintentional fatal <strong>paraquat</strong> poisonings among<br />

agricultural workers in Costa Rica: Report <strong>of</strong> 15 cases. American Journal <strong>of</strong> Industrial Medicine 32:433–441.<br />

Wesseling C, van Wendel de Joode B, Ruepert C, Leon C, Monge P, Hermosillo H, Partanen TJ. 2001. Paraquat<br />

in developing countries. International Journal <strong>of</strong> Occupational and Environmental Health 7:275–286.<br />

Yamada K, Fukushima T. 1993. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. II. Organ specificity <strong>of</strong> <strong>paraquat</strong>stimulated<br />

lipid peroxidation in <strong>the</strong> inner membrane <strong>of</strong> mitochondria. Experiments in Toxicology and Pathology<br />

45:375–380.<br />

Yamaoka K, Nakagawa T, Uno T. 1978. Statistical moments in pharmacokinetics. Journal <strong>of</strong> Pharmacokinetics<br />

and Biopharmacology 6:547–558.<br />

Youngman RJ, Elstner EF. 1981. Oxygen species in <strong>paraquat</strong> toxicity: The crypto-OH radical. FEBS Letters<br />

129:265–268.


____________________________________________________Part II – Original <strong>research</strong><br />

CO 2 /O 2<br />

Thermostatic Bath<br />

Flow and Pressure<br />

Recorder<br />

Water-Jacketed<br />

Reservoir<br />

SUPPLEMENTARY FIGURES<br />

Oxigenator<br />

Jacket<br />

Peristaltic<br />

Pump<br />

Flowmeter<br />

Flex Tubes<br />

(water)<br />

Flex Tubes<br />

(perfusate)<br />

Pressure<br />

Transducer<br />

Flowprobe<br />

Bubble-trap<br />

Aferent<br />

Cannula<br />

Perfusion Medium<br />

Water<br />

Carbogen<br />

(95%O 2 /5%CO 2 )<br />

Ventilator<br />

Eferent<br />

Cannula<br />

Fraction<br />

Collector<br />

Fig. A - Scheme <strong>of</strong> <strong>the</strong> experimental set-up implemented to perform <strong>the</strong> experiments<br />

with <strong>the</strong> isolated rat lung.<br />

Pulmonar trunk<br />

Aorta Artery<br />

Inflow cannula<br />

Outflow<br />

cannula<br />

Fig. B - Schematic description <strong>of</strong> <strong>the</strong> surgical procedure followed to isolate <strong>the</strong> lung. 1.<br />

Outflow cannula inserted in <strong>the</strong> left ventricle. 2. Clamp to fix <strong>the</strong> outflow cannula. 3.<br />

Ligature placed around pulmonary and aortic arteries. 4. Inflow cannula placed just<br />

before <strong>the</strong> bifurcation <strong>of</strong> <strong>the</strong> pulmonar artery.<br />

139


Part I – Original <strong>research</strong>____________________________________________________<br />

140


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER II<br />

Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival case<br />

following intake <strong>of</strong> a potential lethal dose<br />

Reprinted from Pediatric Emergency Care 22: 537-540<br />

Copyright© (2006) with kind permission from Lippincott Williams & Wilkins<br />

141


Part I – Original <strong>research</strong>____________________________________________________<br />

142


Acute Paraquat Poisoning<br />

Report <strong>of</strong> a Survival Case Following Intake <strong>of</strong> a Potential Lethal Dose<br />

Ricardo J. Dinis-Oliveira, PharmD,* António Sarmento, PhD,y Paulo Reis, MD,y Augusta Amaro, MD,y<br />

Fernando Remião, PhD,* Maria L. Bastos, PhD,* and Felix Carvalho, PhD*<br />

Abstract: When properly used, <strong>paraquat</strong> (PQ) is a widely used<br />

bipyridil herbicide with a good safety record. Most cases <strong>of</strong> PQ<br />

poisoning result from intentional ingestion, with death resulting<br />

from hypoxemia secondary to lung fibrosis in moderate to severe<br />

poisonings. With high ingestion volumes (>50 mL <strong>of</strong> a 20% wt/vol<br />

formulation), death results from <strong>multiple</strong> <strong>organ</strong> failure and<br />

cardiovascular collapse within 1 week after <strong>into</strong>xication. The<br />

present report describes a successful clinical case regarding <strong>the</strong><br />

<strong>into</strong>xication <strong>of</strong> a 15-year-old girl by a presumed lethal dose <strong>of</strong> PQ.<br />

The adolescent ingested approximately 50 mL <strong>of</strong> a commercialized<br />

concentrate (20% wt/vol <strong>of</strong> dichloride salt) formulation <strong>of</strong> PQ. High<br />

serum and urinary levels <strong>of</strong> PQ confirmed <strong>the</strong> bad prognosis.<br />

However, <strong>the</strong> <strong>the</strong>rapeutic protocol followed in <strong>the</strong> present clinical<br />

case led to a positive outcome. Besides <strong>the</strong> measures for decreasing<br />

PQ absorption and increasing its elimination, o<strong>the</strong>r protective<br />

procedures were applied in aiming to reduce <strong>the</strong> production <strong>of</strong><br />

reactive oxygen species (ROS), to scavenge ROS, to repair ROS<strong>induced</strong><br />

lesions, and to reduce inflammation. The status-<strong>of</strong>-<strong>the</strong>-art<br />

concerning <strong>the</strong> biochemical and toxicological aspects <strong>of</strong> PQ<br />

poisoning and <strong>the</strong> pharmacologic basis <strong>of</strong> <strong>the</strong> respective treatment<br />

is also presented.<br />

Key Words: <strong>paraquat</strong> poisoning, oral ingestion, lung toxicity<br />

S ince<br />

its introduction in agriculture in 1962, 1<br />

<strong>the</strong><br />

widespread nonselective contact herbicide <strong>paraquat</strong><br />

(PQ), used as desiccant and defoliant in a variety <strong>of</strong> crops,<br />

has caused thousands <strong>of</strong> deaths from both accidental and<br />

voluntary ingestion, as well as from dermal exposure. It may<br />

be considered as one <strong>of</strong> <strong>the</strong> most toxic poisons frequently used<br />

for suicide attempts. A large oral dose <strong>of</strong> PQ (>30 mg kg 1<br />

in humans) rapidly leads to death from multi<strong>organ</strong> failure,<br />

with lung damage consisting <strong>of</strong> disruption <strong>of</strong> alveolar<br />

epi<strong>the</strong>lial cells, hemorrhage, edema, and infiltration <strong>of</strong> inflam-<br />

*REQUIMTE, Department <strong>of</strong> Toxicology, Faculty <strong>of</strong> Pharmacy, University<br />

<strong>of</strong> Porto, Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal and<br />

yIntensive Care Department, Pedro Hispano Hospital, Rua Dr Eduardo<br />

Torres, 4454-509 Matosinhos, Portugal.<br />

Address correspondence and reprint requests to Ricardo Jorge Dinis-<br />

Oliveira, PharmD, and Félix Dias Carvalho, PhD, REQUIMTE,<br />

Department <strong>of</strong> Toxicology, Faculty <strong>of</strong> Pharmacy, University <strong>of</strong> Porto,<br />

Rua Aníbal Cunha, 00351 222003977164, 4099-030 Porto, Portugal.<br />

E-mail: ricardinis@ff.up.pt, felixdc@ff.up.pt.<br />

Copyright n 2006 by Lippincott Williams & Wilkins<br />

ISSN: 0749-5161/06/2207-0537<br />

Pediatric Hospitalist<br />

matory cells <strong>into</strong> <strong>the</strong> interstitial and alveolar spaces. 1 Smaller<br />

doses <strong>of</strong> PQ (from 16 mg kg 1 )mayalsoleadtodeath,butthis<br />

occurs after several days as a result <strong>of</strong> a progressive lung<br />

fibrosis, by proliferation <strong>of</strong> fibroblasts, and excessive collagen<br />

deposition, showing that <strong>the</strong> main target <strong>organ</strong> for PQ toxicity is<br />

<strong>the</strong> lung. 1 The direct cellular toxicity <strong>of</strong> PQ is essentially due to<br />

its redox cycle (Fig. 1): PQ is reduced enzymatically, mainly by<br />

<strong>the</strong> reduced form <strong>of</strong> nicotinamide adenine dinucleotide<br />

phosphate–cytochrome P-450 reductase, <strong>the</strong> reduced form <strong>of</strong><br />

nicotinamide adenine dinucleotide phosphate–cytochrome c<br />

reductase, and <strong>the</strong> reduced form <strong>of</strong> nicotinamide adenine<br />

dinucleotide:ubiquinone oxidoreductase (complex I), to form a<br />

PQ monocation free radical. The PQ monocation free radical is<br />

<strong>the</strong>n rapidly reoxidized in <strong>the</strong> presence <strong>of</strong> oxygen, thus resulting<br />

in <strong>the</strong> generation <strong>of</strong> <strong>the</strong> superoxide radical. 1 This <strong>the</strong>n sets in <strong>the</strong><br />

well-known cascade, leading to generation <strong>of</strong> <strong>the</strong> hydroxyl<br />

radical and consequent deleterious effects. Nowadays, no<br />

antidote or efficient treatment <strong>of</strong> PQ poisoning has been<br />

identified, <strong>the</strong> survival depending on <strong>the</strong> amount ingested and<br />

<strong>the</strong> time elapsed until <strong>the</strong> patient is submitted to intensive<br />

medical measures to inactivate and eliminate PQ.<br />

The aim <strong>of</strong> this article is to report a successful clinical<br />

case regarding <strong>the</strong> <strong>into</strong>xication <strong>of</strong> a young girl who ingested<br />

a potentially lethal dose <strong>of</strong> PQ.<br />

CASE<br />

A 15-year-old girl voluntarily ingested approximately 50 mL<br />

<strong>of</strong> a commercialized PQ formulation (20% wt/vol <strong>of</strong> PQ dichloride<br />

salt), corresponding to nearly 10g <strong>of</strong> PQ ingested. The weight <strong>of</strong> <strong>the</strong><br />

girl weight was 47 kg. Twenty minutes after ingestion <strong>the</strong><br />

adolescent vomited, <strong>the</strong> gastric contents having a greenish<br />

appearance. She was taken to a local hospital about 2 hours 30<br />

minutes after ingestion. After admission she was immediately<br />

submitted to a gastric lavage with physiologic 0.9% NaCl solution.<br />

Mineral adsorbent (100 g <strong>of</strong> Fuller earth) was subsequently given to<br />

reduce fur<strong>the</strong>r absorption <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> bloodstream. The patient<br />

was <strong>the</strong>n transferred to <strong>the</strong> intensive care unit. She was conscious,<br />

anxious, with a coherent speech, presenting a slightly sinus<br />

tachycardia (around 110 heartbeats min 1 ) and a respiratory rate<br />

approximately 22 cycles min 1 , with no fever and hemodynamically<br />

stable. There was no history <strong>of</strong> respiratory or o<strong>the</strong>r illness.<br />

Chest radiograph, hemogram, and blood chemistry were all normal.<br />

No erosion lesions were noted in <strong>the</strong> oral cavity. The remainder <strong>of</strong><br />

physical examination was unremarkable. The ingestion was<br />

Pediatric Emergency Care Volume 22, Number 7, July 2006 537<br />

Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction <strong>of</strong> this article is prohibited.


Dinis-Oliveira et al Pediatric Emergency Care Volume 22, Number 7, July 2006<br />

FIGURE 1. Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> PQ toxicity. A indicates cellular diaphorases. CAT, catalase; FR, Fenton<br />

reaction; Gred, glutathione reductase; GPX, glutathione peroxidase; HWR, Haber-Weiss reaction; PQ + , PQ cation radical; PQ 2+ ,<br />

<strong>paraquat</strong>; SOD, superoxide dismutase or spontaneously.<br />

confirmed by a qualitative sodium dithionite test on a urine sample.<br />

The initial urine colorimetric test showed a dark blue color.<br />

Analysis <strong>of</strong> urine samples collected at 4, 6, 10, 16, 20, 26, 30,<br />

42, and 50 hours after ingestion revealed values <strong>of</strong> 102.83<br />

(prehemoperfusion), 87.07 (during hemoperfusion), 11.97 (after<br />

hemoperfusion), 22.99 (prehemoperfusion), 9.45 (after hemoperfusion),<br />

5.75 (prehemoperfusion), 0.25 (after hemoperfusion), 0.35<br />

(prehemoperfusion), and 0.20 mg L 1 (after hemoperfusion),<br />

respectively. Fifty-eight hours after ingestion, PQ in <strong>the</strong> urine was<br />

lower than <strong>the</strong> quantification limit. The PQ levels in <strong>the</strong> serum<br />

samples at <strong>the</strong> same times were lower than <strong>the</strong> quantification limit<br />

<strong>of</strong> <strong>the</strong> method, with <strong>the</strong> exception <strong>of</strong> <strong>the</strong> first, second, and third<br />

sampled times where it was found (3.5, 1.75, and 0.95 mg L 1 <strong>of</strong><br />

PQ, respectively). Serum and urinary levels <strong>of</strong> PQ were undetectable<br />

20 and 72 hours after ingestion, respectively.<br />

In <strong>the</strong> face <strong>of</strong> <strong>the</strong> severity <strong>of</strong> <strong>the</strong> <strong>into</strong>xication, with high serum<br />

and urine PQ levels, it was followed an aggressive <strong>the</strong>rapeutic protocol.<br />

The patient was submitted to hemoperfusion during 4 days, in 7<br />

sessions <strong>of</strong> 3 hours each. The first session was initiated 4 hours after<br />

ingestion. The observed complications were electrolyte disturbances,<br />

with hypokalemia, hypomagnesemia, and hypophosphatemia. Severe<br />

alterations in <strong>the</strong> coagulation tests due to <strong>the</strong> heparin used during <strong>the</strong><br />

hemoperfusion did not require immediate correction. Only after <strong>the</strong> last<br />

session was protamine sulfate needed. Thrombocytopenia evolved, but<br />

it was resolved after suspension <strong>of</strong> hemoperfusion.<br />

Pharmaco<strong>the</strong>rapy was initiated with (1) 15 mg kg 1<br />

cyclophosphamide (CP) in 100 mL <strong>of</strong> a 5% dextrose solution<br />

perfused over 60 minutes once daily after hemoperfusion during<br />

<strong>the</strong> first 2 days <strong>of</strong> hospitalization; (2) 15 mg kg 1 methylprednisolone<br />

(MP) in 200 mL <strong>of</strong> a 5% dextrose solution perfused over 60<br />

minutes and repeated once daily for 3 consecutive days always after<br />

hemoperfusion; (3) 100 mg kg 1 desferrioxamine (DFO) in 500 mL<br />

<strong>of</strong> a 5% dextrose solution in continuous intravenous perfusion at<br />

21 mL hour 1 during 24 hours in 1 administration started after <strong>the</strong><br />

first hemoperfusion session; (4) 300 vitamin E mg/p.o. twice daily<br />

after hemoperfusion; (5) N-acetylcysteine (NAC) was administered<br />

after <strong>the</strong> first hemoperfusion session in a dose <strong>of</strong> 150 mg kg 1 in<br />

500 mL <strong>of</strong> a 5% dextrose solution perfused during 3 hours;<br />

subsequently, it was given 300 mg kg 1 in 500 mL <strong>of</strong> a 5% dextrose<br />

solution in continuous perfusion at 21 mL hour 1 during 3 weeks.<br />

After 3 days, MP was suspended, and <strong>the</strong> patient received<br />

5 mg <strong>of</strong> intravenous dexamethasone (DX) every 8 hours during <strong>the</strong><br />

next 5 days, with posterior withdraw <strong>the</strong>rapy regime for<br />

approximately 20 days. In addition, <strong>the</strong> patient received prophylaxis<br />

for stress ulcer (40 mg omeprazol, i.v., twice daily) and for<br />

opportunistic infections (one tablet daily containing 800 mg<br />

cotrimoxazol and 160 mg <strong>of</strong> trimethoprim). The patient did not<br />

develop renal or hepatic failure. No signs <strong>of</strong> infection were noted<br />

from long-term steroid <strong>the</strong>rapy. Initial pulmonary function tests and<br />

chest radiograph at <strong>the</strong> time <strong>of</strong> admission were normal. However,<br />

on day 7, computerized axial tomography (CAT) <strong>of</strong> <strong>the</strong> thorax<br />

revealed areas <strong>of</strong> pulmonary densification, with ground-glass<br />

attenuation at <strong>the</strong> lung base possibly indicating initial edema as a<br />

sign <strong>of</strong> fibrosis. Pulmonary function tests also showed alterations in<br />

<strong>the</strong> CO diffusion. The hospitalization lasted 22 days. At discharge<br />

time CO diffusion test and CAT were all normal. Six months later,<br />

all <strong>the</strong> parameters were standard.<br />

DISCUSSION<br />

In <strong>the</strong> present report a successful clinical case is<br />

presented regarding <strong>the</strong> <strong>into</strong>xication <strong>of</strong> a young girl by a<br />

presumed lethal dose <strong>of</strong> PQ.<br />

538 n 2006 Lippincott Williams & Wilkins<br />

Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction <strong>of</strong> this article is prohibited.


Pediatric Emergency Care Volume 22, Number 7, July 2006 Surviving to Paraquat<br />

Previous works 2,3 reporting human PQ poisoning show<br />

that <strong>the</strong> plasma and urine concentration within <strong>the</strong> first 24 and<br />

48 hours post<strong>into</strong>xication are good predictors <strong>of</strong> outcome.<br />

The initial urine colorimetric test showed dark blue color. The<br />

appearance <strong>of</strong> strongly positive dithionite tests in urine and<br />

<strong>the</strong> PQ serum concentration 4, 6, and 10 hours after<br />

ingestion 2,3 were indicators <strong>of</strong> a poor prognosis. 3 Despite<br />

undetectable levels <strong>of</strong> PQ in <strong>the</strong> serum 10 hours after<br />

ingestion, <strong>the</strong> very high concentration <strong>of</strong> PQ in urine was<br />

certainly predictive <strong>of</strong> a fatal outcome. 3 According to survival<br />

probability using <strong>the</strong> criteria <strong>of</strong> <strong>the</strong> study by Scherrmann<br />

et al, 3 Proudfoot et al, 2 and Hart et al, 4 <strong>the</strong> likelihood <strong>of</strong><br />

mortality falls <strong>into</strong> death from pulmonary fibrosis.<br />

Because <strong>the</strong>re are no known antidotes for PQ and <strong>the</strong>re<br />

are no chelating agents capable <strong>of</strong> binding <strong>the</strong> PQ in <strong>the</strong><br />

blood or o<strong>the</strong>r tissues, over <strong>the</strong> past 40 years, strategies in<br />

<strong>the</strong> management <strong>of</strong> PQ poisoning have been directed toward<br />

<strong>the</strong> modification <strong>of</strong> <strong>the</strong> toxicokinetics <strong>of</strong> <strong>the</strong> poison by ei<strong>the</strong>r<br />

decreasing its absorption 5 or enhancing its elimination. Such<br />

approaches are intended to prevent <strong>the</strong> accumulation <strong>of</strong> PQ<br />

in tissues and include procedures such as <strong>induced</strong> emesis or<br />

diarrhea, gastric lavage, administration <strong>of</strong> oral absorbents,<br />

hemodialysis, and hemoperfusion. 6 No vomit induction was<br />

performed because <strong>the</strong> formulation already contained an<br />

emetic, and <strong>the</strong> young girl vomited, which certainly<br />

contributed to <strong>the</strong> positive outcome despite <strong>the</strong> high quantities<br />

<strong>of</strong> PQ that were absorbed and quantified in <strong>the</strong> serum and<br />

urine. After admission, she was immediately submitted to a<br />

gastric lavage with physiologic 0.9% NaCl solution, a<br />

successful measure in some cases <strong>of</strong> heavy PQ poisoning. 6<br />

Mineral adsorbent (100 g <strong>of</strong> Fuller earth) was subsequently<br />

given to reduce fur<strong>the</strong>r absorption <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> bloodstream.<br />

The supporting references for this <strong>the</strong>rapeutic measure not<br />

only include in vitro 7 and in vivo 8 studies demonstrating <strong>the</strong><br />

strong and tight binding <strong>of</strong> PQ to this adsorbent but also some<br />

successful cases <strong>of</strong> PQ poisoning treatment. 9 The patient was<br />

<strong>the</strong>n submitted to hemoperfusion, which seems to be an<br />

indispensable treatment for patients with acute PQ poisoning,<br />

10 increasing <strong>the</strong> chance <strong>of</strong> survival if started early within 4<br />

hours after ingestion and showing higher extraction ratios 11<br />

for PQ when compared to hemodialysis.<br />

Beside <strong>the</strong>se treatments, additional protective measures<br />

were also adopted: (1) those aimed to prevent <strong>the</strong><br />

generation <strong>of</strong> reactive oxygen species, namely, <strong>the</strong> effective<br />

control <strong>of</strong> iron distribution by DFO; (2) those aimed to<br />

scavenge reactive oxygen species (ROS), including <strong>the</strong><br />

maintenance <strong>of</strong> effective levels <strong>of</strong> antioxidants such as<br />

vitamin E; (3) those aimed to repair <strong>the</strong> ROS-<strong>induced</strong><br />

lesions, particularly <strong>the</strong> maintenance <strong>of</strong> effective levels <strong>of</strong><br />

glutathione by administrating NAC; and (4) those aimed to<br />

reduce inflammation by DX, MP, CP, and NAC.<br />

Pharmaco<strong>the</strong>rapy was initiated with CP and MP.<br />

Although high doses <strong>of</strong> CP and DX treatments, including<br />

intravenous CP (5 mg kg 1 d 1 ) and DX (24 mg d 1 ) for 14<br />

days have been correlated with 75% survival rate after PQ<br />

poisoning, 12 a subsequent study 13 did not demonstrate <strong>the</strong><br />

usefulness <strong>of</strong> this approach. Therefore, <strong>the</strong> efficacy <strong>of</strong> highdose<br />

CP and DX in PQ poisoning remains controversial.<br />

Recently, a report 14 demonstrated that pulse <strong>the</strong>rapy with CP<br />

and MP might be effective in preventing respiratory failure<br />

and reducing mortality in patients with moderate to severe<br />

PQ poisoning. Pulse <strong>the</strong>rapy with MP is known as a strong<br />

anti-inflammatory treatment in clinical practice, 14 suppressing<br />

ROS production by neutrophils and macrophages, and in<br />

<strong>the</strong> arachidonic acid cascade. 15<br />

Fur<strong>the</strong>rmore, CP exerts a wide range <strong>of</strong> immunomodulatory<br />

effects that influence virtually all components <strong>of</strong> <strong>the</strong><br />

cellular and humoral immune response, and reduce <strong>the</strong> severity<br />

<strong>of</strong> inflammation, 16 <strong>the</strong>refore contributing to <strong>the</strong> overall effect.<br />

In addition, CP-<strong>induced</strong> leukopenia 1 to 2 weeks later may<br />

contribute to reduce pulmonary inflammatory process <strong>of</strong> PQpoisoned<br />

patients. 12<br />

Taking <strong>into</strong> account <strong>the</strong> involvement <strong>of</strong> ROS in <strong>the</strong><br />

toxicity <strong>of</strong> PQ, compounds that can interfere with <strong>the</strong>ir<br />

generation and propagation <strong>of</strong> oxidative stress may be useful<br />

<strong>the</strong>rapeutical tools in <strong>the</strong> treatment <strong>of</strong> PQ poisoning. It has<br />

been shown that DFO can exert its protective effects not only<br />

by iron chelating (and thus inhibiting <strong>the</strong> PQ-<strong>induced</strong><br />

generation <strong>of</strong> hydroxyl radicals) but also by blocking <strong>the</strong><br />

uptake <strong>of</strong> PQ by <strong>the</strong> alveolar type II cells. 17 Concerning <strong>the</strong><br />

use <strong>of</strong> vitamin E (a-tocopherol), this lipid-soluble vitamin<br />

exerts its antioxidant effects by scavenging free radicals and<br />

stabilizing membranes containing polyunsaturated fatty<br />

acids, 18 which may prevent <strong>the</strong> cytotoxic effects <strong>of</strong> PQ.<br />

The use <strong>of</strong> vitamin E is described in several survival cases<br />

after PQ poisoning. 19<br />

NAC has also been used with success in massive<br />

PQ poisoning. 20 NAC, <strong>the</strong> acetylated derivate <strong>of</strong> <strong>the</strong> amino<br />

acid l-cysteine, was administrated because it is an excellent<br />

source <strong>of</strong> sulfhydryl groups. NAC is indeed converted in<br />

<strong>the</strong> body <strong>into</strong> cysteine, <strong>the</strong> rate limiting amino acid for<br />

glutathione syn<strong>the</strong>sis, promoting detoxification and acting<br />

directly as a free radical scavenger. 21 Exposure <strong>of</strong> human<br />

alveolar cells in vitro to PQ has been shown to induce<br />

apoptotic cell death, perhaps via oxidative stress <strong>mechanisms</strong>,<br />

this toxic effect being inhibited by NAC, an effect<br />

attributed to <strong>the</strong> direct scavenging action <strong>of</strong> its sulfhydryl<br />

group. 22<br />

In addition, it was previously shown that<br />

<strong>the</strong> administration <strong>of</strong> NAC to PQ-challenged rats delayed<br />

<strong>the</strong> PQ-<strong>induced</strong> release <strong>of</strong> chemoattractants for neutrophils<br />

in <strong>the</strong> bronchoalveolar lavage fluid and significantly reduced<br />

<strong>the</strong> infiltration <strong>of</strong> inflammatory cells, suggesting<br />

that NAC can also confer its protective effect by delaying<br />

inflammation. 23<br />

At <strong>the</strong> fourth day after <strong>into</strong>xication, 5 mg <strong>of</strong> intravenous<br />

dexamethasone were administered every 8 hours<br />

during <strong>the</strong> next 5 days to prevent <strong>the</strong> inflammation. 24<br />

Prolonged <strong>the</strong>rapy with steroids may increase survival in a<br />

refractory late-stage, but not in an early-stage state <strong>of</strong> adult<br />

respiratory distress syndrome. 25 This effect is attributable to<br />

<strong>the</strong> downregulation <strong>of</strong> circulating macrophages, as well as <strong>of</strong><br />

collagenase activity, and promotion <strong>of</strong> <strong>the</strong> proliferation <strong>of</strong><br />

type II pneumocytes. 25 Therefore, repeated pulse and<br />

continuous steroid <strong>the</strong>rapy may prevent fur<strong>the</strong>r inflammation<br />

and damage <strong>of</strong> pulmonary tissues by superoxide anion in<br />

patients affected by severe PQ poisoning.<br />

n 2006 Lippincott Williams & Wilkins 539<br />

Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction <strong>of</strong> this article is prohibited.


Dinis-Oliveira et al Pediatric Emergency Care Volume 22, Number 7, July 2006<br />

In conclusion, <strong>the</strong> <strong>the</strong>rapeutic protocol followed in <strong>the</strong><br />

present clinical case was coincidental with a positive outcome.<br />

We conducted an intensive and aggressive treatment based<br />

on <strong>the</strong> high ingestion volume, confirmed by <strong>the</strong> high urine<br />

and serum PQ levels. The prognosis <strong>of</strong> this <strong>into</strong>xicated girl<br />

resembles those patients with moderate to severe poisonings,<br />

which, after a morphologically characterized early destructive<br />

phase <strong>of</strong> alveolar type I and type II epi<strong>the</strong>lial cells, develop a<br />

second proliferative phase defined by alveolitis, pulmonary<br />

edema, and infiltration <strong>of</strong> inflammatory cells. 1 For this reason<br />

our protocol may not be applied to fulminant <strong>into</strong>xications<br />

where multi<strong>organ</strong> failure is <strong>the</strong> main cause <strong>of</strong> death. It is<br />

hoped that <strong>the</strong> present <strong>the</strong>rapeutic approach may be valuable<br />

for o<strong>the</strong>r intensive care units in <strong>the</strong> management <strong>of</strong> this very<br />

common <strong>into</strong>xication. Never<strong>the</strong>less, fur<strong>the</strong>r controlled studies<br />

are required to confirm <strong>the</strong> usefulness <strong>of</strong> our protocol.<br />

REFERENCES<br />

1. Onyeama HP, Oehme FW. A literature review <strong>of</strong> <strong>paraquat</strong> toxicity. Vet<br />

Hum Toxicol. 1984;26:494–502.<br />

2. Proudfoot AT, Stewart MS, Levitt T, et al. Paraquat poisoning:<br />

significance <strong>of</strong> plasma-<strong>paraquat</strong> concentrations. Lancet. 1979;18:<br />

330–332.<br />

3. Scherrmann JM, Houze P, Bismuth C, et al. Prognostic value <strong>of</strong> plasma<br />

and urine <strong>paraquat</strong> concentration. Hum Toxicol. 1987;6:91–93.<br />

4. Hart TB, Nevitt A, Whitehead A. A new statistical approach to<br />

<strong>the</strong> prognostic significance <strong>of</strong> plasma concentrations. Lancet. 1984;2:<br />

1222–1223.<br />

5. Okonek S, H<strong>of</strong>mann A, Henningsen B. Efficacy <strong>of</strong> gut lavage,<br />

hemodialysis, and hemoperfusion in <strong>the</strong> <strong>the</strong>rapy <strong>of</strong> <strong>paraquat</strong> or diquat<br />

<strong>into</strong>xication. Arch Toxicol. 1976;36:43–51.<br />

6. Meredith TJ, Vale JA. Treatment <strong>of</strong> <strong>paraquat</strong> poisoning in man: methods<br />

to prevent absorption. Hum Toxicol. 1987;6:49–55.<br />

7. Okonek S, Setyadharma H, Borchert A, et al. Activated charcoal is as<br />

effective as Fuller’s earth or bentonite in <strong>paraquat</strong> poisoning. Klin<br />

Wochenschr. 1982;60:207–210.<br />

8. Idid SZ, Lee CY. Effects <strong>of</strong> Fuller’s earth and activated charcoal on oral<br />

absorption <strong>of</strong> <strong>paraquat</strong> in rabbits. Clin Exp Pharmacol Physiol.<br />

1996;23:679–681.<br />

9. Clark DG. Inhibition <strong>of</strong> <strong>the</strong> absorption <strong>of</strong> <strong>paraquat</strong> from <strong>the</strong> gastrointestinal<br />

tract by adsorbents. Br J Ind Med. 1971;28:186–188.<br />

10. Hong SF, Yang JO, Lee EY, et al. Effect <strong>of</strong> haemoperfusion on plasma<br />

<strong>paraquat</strong> concentration in vitro and in vivo. Toxicol Ind Health.<br />

2003;17–23.<br />

11. Winchester JF. Dialysis and hemoperfusion in poisoning. Adv Ren<br />

Replace Ther. 2002;9:26–30.<br />

12. Addo E, Poon-King T. Leukocyte suppression in treatment <strong>of</strong> 72<br />

patients with <strong>paraquat</strong> poisoning. Lancet. 1986;1:1117–1120.<br />

13. Perriens JH, Benimadho S, Kiauw IL, et al. High-dose cyclophosphamide<br />

and dexamethasone in <strong>paraquat</strong> poisoning: a prospective study.<br />

Hum Exp Toxicol. 1992;11:129–134.<br />

14. Lin JL, Wei MC, Liu YC. Pulse <strong>the</strong>rapy with cyclophosphamide and<br />

methylprednisolone in patients with moderate to severe <strong>paraquat</strong><br />

poisoning: a preliminary report. Thorax. 1996;51:661–663.<br />

15. Yoshida T, Tanaka M, Sotomatsu A, et al. Effect <strong>of</strong> methylprednisolonepulse<br />

<strong>the</strong>rapy on superoxide production <strong>of</strong> neutrophils. Neurol Res.<br />

1999;21:509–512.<br />

16. Fox DA, McCune WJ. Immunosuppressive drug <strong>the</strong>rapy <strong>of</strong> systemic<br />

lupus ery<strong>the</strong>matosus. Rheum Dis Clin North Am. 1994;20:265–299.<br />

17. Wal NAVd, Smith LL, Oirshot JFv, et al. Effect <strong>of</strong> iron chelators on<br />

<strong>paraquat</strong> toxicity in rats and alveolar type II cells. Am Rev Respir Dis.<br />

1992;145:180–186.<br />

18. Burton GW. Vitamin E: molecular and biological function. Proc Nutr<br />

Soc. 1994;53:251–262.<br />

19. Shahar E, Barzilay Z, Aladjem M. Paraquat poisoning in a child: vitamin<br />

E in amelioration <strong>of</strong> lung injury. Arch Dis Child. 1980;55:830–831.<br />

20. Drault JN, Baelen E, Mehdaoui H, et al. Massive <strong>paraquat</strong> poisoning.<br />

Favorable course after treatment with n-acetylcysteine and early<br />

hemodialysis. Ann Fr Anesth Reanim. 1999;18:534–537.<br />

21. Moldeus P, Cotgreave IA, Berggren M. Lung protection by a thiolcontaining<br />

antioxidant: N-acetylcysteine. Respiration. 1986;50:31–42.<br />

22. Cappelletti G, Maggioni MG, Maci R. Apoptosis in human lung<br />

epi<strong>the</strong>lial cells: triggering by <strong>paraquat</strong> and modulation by antioxidants.<br />

Cell Biol Int. 1998;22:671–678.<br />

23. H<strong>of</strong>fer E, Avidor I, Benjaminov O, et al. N-acetylcysteine delays <strong>the</strong><br />

infiltration <strong>of</strong> inflammatory cells <strong>into</strong> <strong>the</strong> lungs <strong>of</strong> <strong>paraquat</strong>-<strong>into</strong>xicated<br />

rats. Toxicol Appl Pharmacol. 1993;120:8–12.<br />

24. Chen GH, Lin JL, Huang YK. Combined methylprednisolone and<br />

dexamethasone <strong>the</strong>rapy for <strong>paraquat</strong> poisoning. Crit Care Med.<br />

2002;30:2584–2587.<br />

25. Meduri GU, Belenchia JM, Estes RJ, et al. Fibroproliferative phase <strong>of</strong><br />

ARDS. Clinical findings and effects <strong>of</strong> corticosteroids. Chest. 1991;<br />

100:943–952.<br />

540 n 2006 Lippincott Williams & Wilkins<br />

Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction <strong>of</strong> this article is prohibited.


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER III<br />

P-glycoprotein induction: an antidotal pathway for<br />

<strong>paraquat</strong>-<strong>induced</strong> lung toxicity<br />

Reprinted from Free Radical Biology & Medicine 41: 1213–1224<br />

Copyright© (2006) with kind permission from Elsevier Science Inc<br />

147


Part II – Original <strong>research</strong>____________________________________________________<br />

148


Original Contribution<br />

P-glycoprotein induction: an antidotal pathway for<br />

<strong>paraquat</strong>-<strong>induced</strong> lung toxicity ☆<br />

R.J. Dinis-Oliveira a,⁎ , F. Remião a , J.A. Duarte b , R. Ferreira b , A. Sánchez Navarro c ,<br />

M.L. Bastos a , F. Carvalho a,⁎<br />

a REQUIMTE, Department <strong>of</strong> Toxicology, Faculty <strong>of</strong> Pharmacy, University <strong>of</strong> Porto, Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal<br />

b Department <strong>of</strong> Sport Biology, Faculty <strong>of</strong> Sport Sciences, University <strong>of</strong> Porto, Rua Dr. Plácido Costa, 91, 4200-450 Porto, Portugal<br />

c Department <strong>of</strong> Pharmacy and Pharmaceutical Technology, Faculty <strong>of</strong> Pharmacy, University <strong>of</strong> Salamanca, Avenida Campo Charro s/n, 37007 Salamanca, Spain<br />

Abstract<br />

Received 6 April 2006; revised 26 June 2006; accepted 27 June 2006<br />

Available online 3 July 2006<br />

The widespread use <strong>of</strong> <strong>the</strong> nonselective contact herbicide <strong>paraquat</strong> (PQ) has been <strong>the</strong> cause <strong>of</strong> thousands <strong>of</strong> deaths from both accidental and<br />

voluntary ingestion. The main target <strong>organ</strong> for PQ toxicity is <strong>the</strong> lung. No antidote or effective treatment to decrease PQ accumulation in <strong>the</strong> lung<br />

or to disrupt its toxicity has yet been developed. The present study describes a procedure that leads to a remarkable decrease in PQ accumulation in<br />

<strong>the</strong> lung, toge<strong>the</strong>r with an increase in its fecal excretion and a subsequent decrease in several biochemical and histopathological biomarkers <strong>of</strong><br />

toxicity. The administration <strong>of</strong> dexamethasone (100 mg/kg ip) to Wistar rats, 2 h after PQ <strong>into</strong>xication (25 mg/kg ip), decreased <strong>the</strong> lung PQ<br />

accumulation to about 40% <strong>of</strong> <strong>the</strong> group exposed to only PQ and led to an improvement in tissue healing in just 24 h as a result <strong>of</strong> <strong>the</strong> induction <strong>of</strong><br />

de novo syn<strong>the</strong>sis <strong>of</strong> P-glycoprotein (P-gp). The involvement <strong>of</strong> P-gp in <strong>the</strong>se effects was confirmed by Western blot analysis and by <strong>the</strong> use <strong>of</strong> a<br />

competitive inhibitor <strong>of</strong> this transporter, verapamil (10 mg/kg ip), which, given 1 h before dexamethasone, blocked its protective effects, causing<br />

instead an increase in lung PQ concentration and an aggravation <strong>of</strong> toxicity. In conclusion, <strong>the</strong> induction <strong>of</strong> P-gp, leading to a decrease in lung<br />

levels <strong>of</strong> PQ and <strong>the</strong> consequent prevention <strong>of</strong> toxicity, seems to be a new and promising treatment for PQ poisonings that should be fur<strong>the</strong>r<br />

clinically tested.<br />

© 2006 Elsevier Inc. All rights reserved.<br />

Keywords: Paraquat; Lung toxicity; P-glycoprotein; Dexamethasone; Free radicals<br />

Paraquat dichloride (methyl viologen; PQ) is an effective and<br />

widely used herbicide as desiccant and defoliant in a variety <strong>of</strong><br />

crops. Despite PQ being <strong>the</strong> third most extensively used<br />

herbicide in <strong>the</strong> world, it can be considered one <strong>of</strong> <strong>the</strong> most toxic<br />

over <strong>the</strong> past 60 years. Indeed, PQ has caused thousands <strong>of</strong><br />

deaths from both accidental and voluntary ingestion, as well as<br />

from dermal exposure [1]. Depending on <strong>the</strong> ingested dose,<br />

different clinical patterns and outcomes have been observed in<br />

animals and humans [1]. A large oral dose <strong>of</strong> PQ (>30 mg/kg in<br />

humans) rapidly leads to death from multi<strong>organ</strong> failure, with<br />

☆ Portuguese Patent Pending 103420.<br />

⁎ Corresponding author. Fax: +351222003977.<br />

E-mail addresses: ricardinis@ff.up.pt (R.J. Dinis-Oliveira),<br />

felixdc@ff.up.pt (F. Carvalho).<br />

0891-5849/$ - see front matter © 2006 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.freeradbiomed.2006.06.012<br />

Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

www.elsevier.com/locate/freeradbiomed<br />

lung damage consisting <strong>of</strong> disruption <strong>of</strong> alveolar epi<strong>the</strong>lial cells<br />

(type I and II pneumocytes) and bronchiolar Clara cells,<br />

hemorrhage, edema, hypoxemia, and infiltration <strong>of</strong> inflammatory<br />

cells <strong>into</strong> <strong>the</strong> interstitial and alveolar spaces [1]. Smaller<br />

doses <strong>of</strong> PQ (from 16 mg/kg) may also lead to death, but this<br />

occurs after several days as a result <strong>of</strong> a progressive lung<br />

fibrosis and consequent respiratory failure, by proliferation <strong>of</strong><br />

fibroblasts and excessive collagen deposition [1].<br />

In 1974, Rose et al. [2] demonstrated that <strong>the</strong> accumulation<br />

<strong>of</strong> radioactively labeled [ 14 C]PQ in rat lung slices was energy<br />

dependent and obeyed saturation kinetics. O<strong>the</strong>r studies led to<br />

<strong>the</strong> conclusion that PQ accumulated in <strong>the</strong> lung through a<br />

system in which polyamines are <strong>the</strong> natural substrates and that<br />

in comparison to o<strong>the</strong>r <strong>organ</strong>s, <strong>the</strong> lungs, and more specifically<br />

<strong>the</strong> alveolar epi<strong>the</strong>lial and Clara cells, were endowed with a<br />

particularly active polyamine uptake system [3]. Although PQ


1214 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

proved to be a “poor” substrate for <strong>the</strong> polyamine uptake<br />

system, it undoubtedly accumulates in <strong>the</strong> lung through this<br />

transport pathway. The mechanism <strong>of</strong> PQ-<strong>induced</strong> acute lung<br />

toxicity is well known. It is essentially due to its redox cycle<br />

[4]: PQ is reduced enzymatically, mainly by NADPH-cytochrome<br />

P450 reductase [5] and NADH:ubiquinone oxidoreductase<br />

(complex I) [6–8], to form a PQ monocation free<br />

radical. The PQ monocation free radical is <strong>the</strong>n rapidly<br />

reoxidized in <strong>the</strong> presence <strong>of</strong> oxygen (which has a high partial<br />

pressure in lungs) with <strong>the</strong> subsequent generation <strong>of</strong> <strong>the</strong><br />

S− superoxide radical (O2 ) [9,10]. This <strong>the</strong>n begins <strong>the</strong> wellknown<br />

cascade leading to <strong>the</strong> production <strong>of</strong> o<strong>the</strong>r reactive<br />

oxygen species (ROS), mainly hydrogen peroxide (H2O2) and<br />

hydroxyl radical (HOS ).<br />

Currently, no antidote or effective treatment for PQ<br />

poisoning has been identified, survival being mainly dependent<br />

on <strong>the</strong> amount ingested and <strong>the</strong> time elapsed until <strong>the</strong> patient is<br />

submitted to intensive medical measures to inactivate or to<br />

eliminate PQ, before its cellular uptake. These approaches<br />

include procedures such as induction <strong>of</strong> emesis or intestinal<br />

transit, gastric lavage, administration <strong>of</strong> oral adsorbents,<br />

hemodialysis, and hemoperfusion [11–13]. In addition to<br />

<strong>the</strong>se treatments, protective measures have also been adopted:<br />

(i) to prevent <strong>the</strong> generation <strong>of</strong> ROS, namely <strong>the</strong> effective iron<br />

chelation by desferrioxamine [14]; (ii) to scavenge ROS,<br />

including <strong>the</strong> maintenance <strong>of</strong> effective levels <strong>of</strong> antioxidants<br />

[15]; and (iii) to reduce <strong>the</strong> inflammation [16,17]. However,<br />

such treatments have a general low efficacy and <strong>the</strong> fatality rate<br />

remains very high. It was precisely this lacuna in <strong>the</strong> treatment<br />

<strong>of</strong> PQ <strong>into</strong>xication that impelled our study.<br />

In this work we propose a new approach for PQ poisonings<br />

by induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> a plasma membrane<br />

phosphoglycoprotein, P-glycoprotein (P-gp). P-gp, a member <strong>of</strong><br />

<strong>the</strong> ATP-binding cassette superfamily, was initially identified in<br />

tumor cells as an ATP-dependent transporter, which can export<br />

a wide variety <strong>of</strong> unmodified substrates out <strong>of</strong> <strong>the</strong> cell, namely<br />

Vinca alkaloids, colchicine, antibiotics, anthracyclines, cardiac<br />

glycosides, <strong>organ</strong>ic cations, and pesticides [18–20]. This drug<br />

transport occurs against <strong>the</strong> concentration gradient and is<br />

independent <strong>of</strong> an electrochemical transmembrane potential or<br />

proton gradient [21].<br />

A number <strong>of</strong> reports exist noting that dexamethasone (DEX)<br />

induces P-gp levels in liver, brain, and intestinal tissue and also<br />

in lung tissue [22], an effect that seems to be glucocorticoid<br />

concentration-dependent. This induction phenomenon is rapid,<br />

because a maximum effect may be observed 24 h after a single<br />

administration [22]. On <strong>the</strong> o<strong>the</strong>r hand, P-gp-mediated efflux<br />

can be pharmacologically inhibited using several drugs, namely<br />

verapamil (VER), cyclosporin A, and amiodarone [23]. Thus, in<br />

<strong>the</strong> present study, we investigated, for <strong>the</strong> first time, <strong>the</strong> process<br />

<strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-gp, by DEX, in <strong>the</strong> concentration <strong>of</strong><br />

PQ in rat lung and its urinary and fecal excretion. The<br />

preventive effect <strong>of</strong> this pharmacological approach against PQ<strong>induced</strong><br />

lung toxicity was also evaluated using both biochemical<br />

and histopathological biomarkers <strong>of</strong> toxicity. VER, as<br />

competitive inhibitor <strong>of</strong> P-gp, was used to confirm <strong>the</strong> importance<br />

<strong>of</strong> this transporter in PQ excretion.<br />

Materials and methods<br />

Chemicals and drugs<br />

Paraquat dichloride (1,1′-dimethyl-4,4′-bipyridinium dichloride),<br />

dexamethasone [(11β,16α)-9-fluoro-11,17,21-trihydroxy-16-methylpregna-1,4-diene-3,20-dione],<br />

(±)-verapamil<br />

hydrochloride (5-[(3,4-dimethoxyphenethyl)methylamino]-2-<br />

(3,4-dimethoxyphenyl)-2-isopropylvaleronitrile hydrochloride),<br />

3,3′,5,5′-tetramethylbenzidine (TMB), 5-sulfosalicylic<br />

acid, NADPH (nicotinamide adenine dinucleotide phosphate<br />

reduced), GSH (reduced glutathione), GSSG (oxidized glutathione),<br />

2-vinylpiridine, and 2,4-dinitrophenylhydrazine<br />

(DNPH) were all obtained from Sigma (St. Louis, MO,<br />

USA). Saline solution (NaCl 0.9%) and sodium thiopental<br />

were obtained from B. Braun (Lisbon, Portugal). Sodium<br />

hydroxide (NaOH), sodium dithionite (Na 2S 2O 4), 2-thiobarbituric<br />

acid (C 4H 4N 2O 2S), and trichloroacetic acid (Cl 3CCOOH)<br />

were obtained from Merck (Darmstadt, Germany). All <strong>the</strong><br />

reagents used were <strong>of</strong> analytical grade or <strong>of</strong> <strong>the</strong> highest<br />

available grade.<br />

Animals and experimental design<br />

The study was performed using adult male Wistar rats<br />

obtained from Charles River S.A. (Barcelona, Spain), with a<br />

mean weight <strong>of</strong> 252±8 g. Animals were kept under standard<br />

laboratory conditions (12/12 h light/darkness, 22±2°C room<br />

temperature, 50–60% humidity) for at least 1 week (quarantine)<br />

before starting <strong>the</strong> experiments. Animals were allowed<br />

access to tap water and rat chow ad libitum during <strong>the</strong><br />

quarantine period. Animal experiments were licensed by <strong>the</strong><br />

Portuguese General Directorate <strong>of</strong> Veterinary Medicine.<br />

Housing and experimental treatment <strong>of</strong> animals were in<br />

accordance with <strong>the</strong> Guide for <strong>the</strong> Care and Use <strong>of</strong><br />

Laboratory Animals from <strong>the</strong> Institute for Laboratory Animal<br />

Research (ILAR 1996). The experiments complied with <strong>the</strong><br />

current laws <strong>of</strong> Portugal.<br />

After <strong>the</strong> quarantine period, 52 animals were randomly<br />

divided <strong>into</strong> four groups <strong>of</strong> 13 animals each. Each animal was<br />

individually housed in a metabolic cage where it was kept<br />

during <strong>the</strong> whole time <strong>of</strong> experiment (26 h). Animals were<br />

fasted during <strong>the</strong> entire experimental period but water was given<br />

ad libitum. Urine and feces were collected over ice during <strong>the</strong><br />

26-h period, for quantification <strong>of</strong> PQ.<br />

The administrations <strong>of</strong> vehicle (0.9% NaCl), PQ, DEX, and<br />

VER were all done intraperitoneally (ip) in an injection<br />

volume <strong>of</strong> 0.5 ml. The four groups were treated as follows: (i)<br />

The control group (n=13) animals were treated with 0.9%<br />

NaCl. Animals were treated with two more administrations <strong>of</strong><br />

0.9% NaCl, 1 and 2 h later, respectively. (ii) The PQ group<br />

(n =13) animals were <strong>into</strong>xicated with PQ (25 mg/kg).<br />

Animals were treated with two administrations <strong>of</strong> 0.9%<br />

NaCl, 1 and 2 h later, respectively. (iii) The PQ+DEX group<br />

(n =13) animals were <strong>into</strong>xicated with PQ (25 mg/kg).<br />

Animals were treated with 0.9% NaCl and DEX (100 mg/<br />

kg), 1 and 2 h later, respectively. The schedule <strong>of</strong> DEX


administration was chosen considering <strong>the</strong> lag time necessary<br />

for <strong>the</strong> arrival <strong>of</strong> <strong>the</strong> patient at <strong>the</strong> hospital after PQ <strong>into</strong>xication.<br />

(iv) The PQ+DEX +VER group (n=13) animals were <strong>into</strong>xicated<br />

with PQ (25 mg/kg). Animals were treated with VER<br />

(10 mg/kg) and DEX (100 mg/kg), 1 and 2 h later, respectively.<br />

The experimental dose <strong>of</strong> DEX has been applied in<br />

numerous studies for inducing de novo syn<strong>the</strong>sis <strong>of</strong> P-gp [22].<br />

The VER dose was selected according to some reported studies<br />

referring to P-gp competitive inhibition in vivo [24]. The PQ<br />

dose was similar to that used in previous studies resulting in<br />

severe lung toxicity [25,26].<br />

The treatments for all groups were always conducted<br />

between 8:00 and 10:00 AM.<br />

Collection and processing <strong>of</strong> lung samples<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

Twenty-six hours after PQ administration, anes<strong>the</strong>sia was<br />

<strong>induced</strong> with sodium thiopental (60 mg/kg, ip). Animals were<br />

placed in <strong>the</strong> decubito supino position and tracheotomy and<br />

tracheal cannulation was done, followed by <strong>the</strong> immediate<br />

connection <strong>of</strong> <strong>the</strong> cannula to a mechanical ventilation system<br />

that supplied a tidal volume <strong>of</strong> 2 ml at a respiratory frequency <strong>of</strong><br />

60 breaths/min. The thorax was opened by two lateral<br />

transversal incisions and one central longitudinal incision to<br />

expose <strong>the</strong> pulmonary artery. In 10 rats <strong>of</strong> each group, lungs<br />

were perfused in situ through <strong>the</strong> pulmonary artery with cold<br />

0.9% NaCl for 3 min at a rate <strong>of</strong> 10 ml/min to be completely<br />

cleaned <strong>of</strong> blood. At <strong>the</strong> same time that this perfusion was<br />

initiated, a cut at <strong>the</strong> left wall ventricle was done to avoid<br />

overpressure. Lungs were removed, cleaned <strong>of</strong> all major<br />

cartilaginous tissues <strong>of</strong> <strong>the</strong> conducting airways, pat-dried with<br />

gauze, weighed, and processed as follows: (i) The right lung<br />

(except <strong>the</strong> posterior lobe) was homogenized (Ultra-Turrax<br />

homogenizer) in a cold mixture <strong>of</strong> phosphate buffer [(KH 2PO 4<br />

+Na 2HPO 4·H 2O) 50 mM, pH 7.4] and 0.1% (v/v) Triton X-<br />

100, 1 g <strong>of</strong> tissue/4 ml <strong>of</strong> mixture, and centrifuged (3000g, 4°C,<br />

for 10 min). Aliquots <strong>of</strong> <strong>the</strong> resulting supernatants were stored<br />

(−80°C) for posterior quantification <strong>of</strong> <strong>the</strong> pulmonary remaining<br />

PQ, myeloperoxidase activity (MPO), superoxide dismutase<br />

activity (SOD), carbonyl groups, and protein levels. Aliquots <strong>of</strong><br />

<strong>the</strong> resulting supernatants were <strong>the</strong>n centrifuged at 33,000g,<br />

4°C, for 30 min. The pellet containing <strong>the</strong> crude membrane<br />

fractions was resuspended in 50 mM mannitol, 20 mM Hepes–<br />

Tris, pH 7.5, and stored at −80°C for posterior confirmation <strong>of</strong><br />

P-gp induction. The posterior lobe was homogenized in<br />

perchloric acid (5% final concentration) and <strong>the</strong>n centrifuged<br />

(13,000g, 4°C, for 10 min). Supernatants were stored (−80°C)<br />

for posterior quantification <strong>of</strong> GSH and GSSG. The pellet was<br />

used for protein quantification. (ii) The left lung was<br />

homogenized (Ultra-Turrax homogenizer) in trichloroacetic<br />

acid 10% (1/4 m/v) and <strong>the</strong>n centrifuged (13,000g, 4°C, for<br />

10 min). Aliquots <strong>of</strong> <strong>the</strong> resulting supernatants were immediately<br />

used for evaluating <strong>the</strong> degree <strong>of</strong> lipid peroxidation (LPO).<br />

The pellet was used for protein quantification.<br />

The relative lung weight (RLW) <strong>of</strong> each animal was<br />

calculated as a percentage <strong>of</strong> <strong>the</strong> absolute body weight on <strong>the</strong><br />

sacrifice day.<br />

Quantification <strong>of</strong> PQ in rat lung, urine, and feces<br />

Aliquots <strong>of</strong> right lung supernatants were treated with 5sulfosalicylic<br />

acid (5% in final volume) and <strong>the</strong>n centrifuged<br />

(13,000g, 4°C, for 10 min).<br />

Feces were treated with 5-sulfosalicylic acid (5% in final<br />

volume) and <strong>the</strong>n centrifuged (13,000g, 4°C, for 20 min). Urine<br />

samples were centrifuged (13,000g, 4°C, for 20 min).<br />

The resulting supernatant fractions from lung, urine, and<br />

feces were alkalinized with 10 N NaOH (pH >9) and <strong>the</strong>n<br />

gently mixed with a few crystals <strong>of</strong> a reductant (sodium<br />

dithionite) to give <strong>the</strong> blue color characteristic <strong>of</strong> <strong>the</strong> PQ cation<br />

radical. PQ quantification was carried out by a previously<br />

reported method based on second-derivative spectrophotometry<br />

[27].<br />

Evaluation <strong>of</strong> P-gp induction<br />

1215<br />

Proteins (40 μg) from lung tissue homogenates <strong>of</strong> <strong>the</strong> four<br />

groups were separated by SDS–PAGE on a 6.25% acrylamide<br />

gel according to <strong>the</strong> method <strong>of</strong> Laemmli [28]. P-gp was<br />

detected by Western blot analysis using <strong>the</strong> polyclonal<br />

antibody sc-1517 (Santa Cruz Biotechnology, Inc.). A horseradish<br />

peroxidase-conjugated anti-goat IgG was used as<br />

secondary antibody. The bands were visualized by treating<br />

<strong>the</strong> immunoblots with ECL chemiluminescence reagents<br />

(Amersham, Pharmacia Biotech, Buckinghamshire, UK),<br />

according to <strong>the</strong> supplier’s instructions, followed by exposure<br />

to X-ray films (Kodak Biomax Light Film; Sigma). The films<br />

were analyzed with QuantityOne S<strong>of</strong>tware (Bio-Rad). Optical<br />

density results were expressed as percentage variation <strong>of</strong><br />

control values.<br />

Tissue processing for structural and ultrastructural qualitative<br />

and semiquantitative analysis<br />

Three animals <strong>of</strong> each group were assigned to histological<br />

analysis. Lung samples were subjected to routine procedures for<br />

light microscopy (LM) and transmission electron microscopy<br />

(TEM) analysis. With <strong>the</strong> animals under anes<strong>the</strong>sia, lung<br />

fixation was initiated in situ by perfusion through <strong>the</strong><br />

pulmonary artery, with 2.5% glutaraldehyde in 0.2 M sodium<br />

cacodylate buffer (pH 7.2–7.4) for 3 min. Subsequently, lungs<br />

were excised, sectioned <strong>into</strong> ∼1-mm 3 pieces and fixed (by<br />

diffusion) in <strong>the</strong> same fixative for 2 h. After two washing steps,<br />

<strong>of</strong> 30 min each with buffer solution, <strong>the</strong> specimens were<br />

dehydrated in graded alcohol for 2 h and <strong>the</strong>n embedded in<br />

Epon. Propylene oxide was <strong>the</strong> compound used in <strong>the</strong><br />

dehydration–impregnation transition. The inclusion phase<br />

lasted 2 days. All <strong>the</strong> procedures were done at 4°C, with <strong>the</strong><br />

exception <strong>of</strong> <strong>the</strong> inclusion phase, which was performed at 60°C.<br />

Subsequent to <strong>the</strong> resin polymerization, semithin sections (1 μm<br />

thick) and ultrathin sections (500 Å thick) were prepared<br />

(Ultracut, Leica), respectively for LM and TEM analysis. The<br />

grids, mounted with <strong>the</strong> ultrathin specimen sections, were<br />

double-contrasted with 0.5% saturated uranyl acetate aqueous<br />

solution for 30 min and <strong>the</strong>n with 0.2% lead citrate solution for


1216 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

15 min. The slides, mounted with semithin sections, were<br />

stained with toluidine blue. Five slides and three grids from<br />

each animal (totaling 10 slides and six grids per group), were<br />

examined in a Zeiss Phomi III photomicroscope and in a<br />

transmission electronic microscope (Zeiss EM 10A).<br />

Histopathological evidence <strong>of</strong> acute tissue damage was<br />

semiquantified according to <strong>the</strong> methodology described<br />

elsewhere [29–32]. For each group, more than 1000 cells<br />

per slide and 100 cells per grid were analyzed in a blind<br />

fashion in order to semiquantify <strong>the</strong> severity and incidence <strong>of</strong><br />

<strong>the</strong> following parameters in every slide or grid: (i) cellular<br />

degeneration, (ii) interstitial inflammatory cell infiltration, (iii)<br />

necrotic zones, and (iv) tissue dis<strong>organ</strong>ization. Considering<br />

<strong>the</strong> cellular degeneration, its severity was scored according to<br />

<strong>the</strong> number <strong>of</strong> cells showing any alterations (dilatation,<br />

vacuolization, pyknotic nuclei, and cellular density) in <strong>the</strong><br />

LM visual field: grade 0, no change from normal; grade 1, a<br />

limited number <strong>of</strong> isolated cells (until 5% <strong>of</strong> <strong>the</strong> total cell<br />

number); grade 2, groups <strong>of</strong> cells (5–30% <strong>of</strong> <strong>the</strong> cell total<br />

number); and grade 3, diffuse cell damage (higher than 30%<br />

<strong>of</strong> <strong>the</strong> total cell number). The severity <strong>of</strong> inflammatory<br />

reaction was scored as grade 0, no cellular infiltration; grade<br />

1, mild leukocyte infiltration (1 to 3 cells by visual field);<br />

grade 2, moderate infiltration (4 to 6 leukocytes by visual<br />

field); and grade 3, heavy infiltration by neutrophils. The<br />

severity <strong>of</strong> necrosis was scored as follows: grade 0, no<br />

necrosis; grade 1, dispersed necrotic foci; grade 2, confluent<br />

necrotic areas; grade 3, massive necrosis. The severity <strong>of</strong><br />

tissue dis<strong>organ</strong>ization was scored according to <strong>the</strong> percentage<br />

<strong>of</strong> <strong>the</strong> affected tissue: grade 0, normal structure; grade 1, less<br />

than one-third <strong>of</strong> tissue; grade 2, greater than one-third and<br />

less than two-thirds; grade 3, greater <strong>of</strong> two-thirds <strong>of</strong> tissue.<br />

For each animal, <strong>the</strong> highest possible total tissue score was<br />

12 and <strong>the</strong> lowest was 0.<br />

An examiner blinded to each tissue sample analyzed all grids<br />

and slides independently.<br />

Protein quantification<br />

Protein quantification was performed according to <strong>the</strong><br />

method <strong>of</strong> Lowry et al. [33] using bovine serum albumin as<br />

standard.<br />

Measurement <strong>of</strong> toxicological biomarkers<br />

LPO was evaluated by <strong>the</strong> thiobarbituric acid-reactive<br />

substances methodology [34]. Results are expressed as nanomoles<br />

<strong>of</strong> malondialdehyde (MDA) equivalents per milligram<br />

protein using an extinction coefficient (ε) <strong>of</strong> 1.56×10 5 M −1<br />

cm −1 .<br />

Protein carbonyl groups (ketones and aldehydes) were<br />

determined according to Levine et al. [35]. Results are<br />

expressed as nanomoles <strong>of</strong> DNPH incorporated per milligram<br />

<strong>of</strong> protein (ε=2.2×10 4 M −1 cm −1 ).<br />

MPO activity was measured according to <strong>the</strong> method<br />

followed by Suzuki et al. and Andrews et al. [36,37], with<br />

slight modifications. Briefly, <strong>the</strong> supernatants were initially<br />

submitted to three cycles <strong>of</strong> snap freezing. The assay mixture<br />

consisted <strong>of</strong> 50 μl <strong>of</strong> supernatant and 50 μl <strong>of</strong> TMB (final<br />

concentration 7.5 mM) dissolved in dimethyl sulfoxide. The<br />

enzymatic activity was initiated by addition <strong>of</strong> 50 μl <strong>of</strong>H2O2<br />

(final concentration 1.5 mM) dissolved in phosphate buffer<br />

(Na2HPO4·2H2O 50 mM, pH 5.4). The rate <strong>of</strong> MPO/H2O2<br />

system-catalyzed oxidation <strong>of</strong> TMB was followed by<br />

recording <strong>the</strong> absorbance increase at 655 nm at 37°C for<br />

3 min. One enzyme unit (U) was defined as <strong>the</strong> amount <strong>of</strong><br />

enzyme capable <strong>of</strong> reducing 1 μl <strong>of</strong>H2O2/min under assay<br />

conditions. Results are expressed in enzyme U/g <strong>of</strong> protein<br />

(ε=3.9×10 4 M − 1 cm − 1 ).<br />

GSH and GSSG concentrations were determined by <strong>the</strong> 5,5′dithiobis-2-nitrobenzoic<br />

acid–GSSG reductase recycling assay<br />

as described before [38]. Results are expressed in nanomoles <strong>of</strong><br />

GSH or GSSG per milligram <strong>of</strong> protein.<br />

Copper/zinc superoxide dismutase (CuZnSOD) and manganese<br />

superoxide dismutase (MnSOD) were assayed using <strong>the</strong><br />

method <strong>of</strong> Flohé and Otting [39] with modifications. A<br />

U−<br />

xanthine–xanthine oxidase system was used to generate O2 .<br />

The subsequent reduction <strong>of</strong> nitroblue tetrazolium (NBT) by<br />

O2<br />

U− was monitored at 560 nm. Potassium cyanide (2 mM) was<br />

used to allow <strong>the</strong> measurement <strong>of</strong> MnSOD. Enzyme activity<br />

was expressed in U/mg <strong>of</strong> protein 1 U <strong>of</strong> SOD is defined as <strong>the</strong><br />

amount <strong>of</strong> enzyme required to inhibit <strong>the</strong> rate <strong>of</strong> NBT reduction<br />

by 50%).<br />

Statistical analysis<br />

Results are expressed as means ±SEM (standard error <strong>of</strong> <strong>the</strong><br />

mean). Statistical comparison between groups was estimated<br />

using <strong>the</strong> nonparametric method <strong>of</strong> Kruskal–Wallis followed by<br />

Dunn’s test. In all cases, p values lower than 0.05 were<br />

considered statistically significant.<br />

Results<br />

Macroscopic observations<br />

Diarrhea, piloerection, weight loss, anorexia, adipsia, hyperpnea,<br />

dyspnea, tachycardia, and a red drainage around <strong>the</strong> mouth,<br />

eyes, and nose were present especially in animals subjected to<br />

only PQ or to PQ+VER+DEX. During <strong>the</strong> experimental period<br />

rats belonging to groups PQ and PQ+VER+DEX did not ingest<br />

any amount <strong>of</strong> water. Deep breathing was observed and <strong>the</strong><br />

thorax was sunken in <strong>the</strong> animals from <strong>the</strong>se groups, in contrast<br />

to those treated with DEX.<br />

Lung PQ concentrations<br />

The concentration <strong>of</strong> PQ in lungs <strong>of</strong> <strong>the</strong> PQ-treated group<br />

was 0.127±0.010 (mean ±SEM; μg/mg protein). Animals<br />

post-treated with DEX evidenced a significant decrease in<br />

PQ lung concentration, down to 0.051±0.012 (p


Urinary and fecal excretion <strong>of</strong> PQ<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

Quantification <strong>of</strong> urinary PQ levels showed that almost all<br />

<strong>the</strong> PQ administered was eliminated by urine within 26 h<br />

(nearly 90%). The inclusion <strong>of</strong> DEX did not result in any<br />

increment <strong>of</strong> urinary excretion <strong>of</strong> PQ. The same result was<br />

obtained in rats exposed to DEX and VER (Fig. 1). On <strong>the</strong><br />

o<strong>the</strong>r hand, and as shown in <strong>the</strong> Fig. 1, rats that received<br />

DEX in addition to PQ (PQ+DEX group) had a significant<br />

increase in PQ fecal excretion, up to 0.651±0.088 (mean±<br />

Fig. 1. Levels <strong>of</strong> PQ in <strong>the</strong> lung, urine, and feces <strong>of</strong> <strong>the</strong> <strong>paraquat</strong> (PQ),<br />

<strong>paraquat</strong>+dexamethasone (PQ+Dex), and <strong>paraquat</strong>+verapamil+dexamethaQ<br />

sone (PQ+Ver+Dex) groups. Values are given as means±SEM (n=10).<br />

ns p>0.05, *p


1218 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

DEX-<strong>induced</strong> P-gp expression<br />

In lung, P-gp expression increased about 1.9-fold (p


cytoplasmic vacuoles, were also identified in <strong>the</strong> interstitial<br />

space (Fig. 4D). Animals from <strong>the</strong> PQ group also revealed an<br />

interstitial edema, indicated by <strong>the</strong> existence <strong>of</strong> intercellular<br />

vacuolization areas that were characterized by a minor density<br />

ultrastructure by TEM. The majority <strong>of</strong> pneumocytes showed, at<br />

least, one ultrastructural abnormality, mitochondrial swelling<br />

being <strong>the</strong> mostly frequent alteration (Fig. 4D). In <strong>the</strong> PQ group,<br />

<strong>the</strong> TEM analysis evidenced a few endo<strong>the</strong>lial cells with<br />

chromatin condensation in <strong>the</strong> nuclear periphery, suggestive <strong>of</strong><br />

<strong>the</strong> occurrence <strong>of</strong> apoptosis, and <strong>the</strong> LM analysis revealed <strong>the</strong><br />

presence <strong>of</strong> several pyknotic nuclei. In <strong>the</strong> PQ+DEX group,<br />

compared to PQ animals, <strong>the</strong> occurrence <strong>of</strong> <strong>the</strong> above-mentioned<br />

alterations was drastically attenuated, particularly <strong>the</strong> amount <strong>of</strong><br />

phagocytes observed in interstitial space or within capillaries<br />

neighboring endo<strong>the</strong>lial cells. Moreover, despite <strong>the</strong> existence <strong>of</strong><br />

several pneumocytes with mitochondrial swelling and evidence<br />

<strong>of</strong> interstitial edema, <strong>the</strong> exuberance <strong>of</strong> those signals<br />

and <strong>the</strong> ratio <strong>of</strong> affected cells were drastically attenuated in<br />

PQ+DEX animals (Figs. 4E and 4F). Fur<strong>the</strong>rmore, compared<br />

to <strong>the</strong> PQ group, <strong>the</strong> vascular congestion and <strong>the</strong> alveolar<br />

collapse were not as evident in <strong>the</strong> PQ+DEX animals (Fig. 4F).<br />

Some pyknotic nuclei were also observed in this group but with<br />

an apparently lower occurrence compared to <strong>the</strong> PQ group. Figs.<br />

4G and 4H illustrate <strong>the</strong> main structural and ultrastructural<br />

alterations observed in animals injected with PQ plus VER and<br />

DEX (PQ+VER+DEX group). In general, <strong>the</strong> majority <strong>of</strong><br />

histological changes detected by LM and TEM had been<br />

qualitatively and quantitatively identical to those observed in <strong>the</strong><br />

PQ group. However, <strong>the</strong> occurrence <strong>of</strong> interstitial phagocyte<br />

infiltration and <strong>the</strong> signals <strong>of</strong> extracellular edema were apparently<br />

more severe in <strong>the</strong> PQ+VER +DEX group.<br />

Concerning <strong>the</strong> semiquantitative analysis, <strong>the</strong> total score<br />

obtained was 0.03±0.02, 1.82±0.09, 0.96±0.08, and 1.90±0.10<br />

in <strong>the</strong> control, PQ, PQ+DEX, and PQ+VER+DEX groups,<br />

respectively. Significant differences were observed between all<br />

groups exposed to PQ and <strong>the</strong> control group (p


1220 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

Fig. 6. Characterization <strong>of</strong> <strong>the</strong> lung antioxidant defenses. GSH (nmol GSH/mg <strong>of</strong> protein) and GSSG (nmol GSSG/mg <strong>of</strong> protein) levels and MnSOD and CuZnSOD<br />

activity (U/mg <strong>of</strong> protein) in <strong>the</strong> control, PQ, PQ+Dex, and PQ+Ver+Dex groups. Values are given as means±SEM (n=10). ns p>0.05, *p


R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

Fig. 7. Proposed scheme for <strong>the</strong> PQ efflux mediated by P-gp. Abbreviations used: GR, glucocorticoid receptor; DNA, deoxyribonucleic acid; ATP, adenosine<br />

triphosphate; ADP, adenosine diphosphate; MDR, multidrug resistance gene.<br />

canalicular membrane <strong>of</strong> hepatocytes [50,51] and PQ excretion<br />

in <strong>the</strong> bile [52] have already reported. Interestingly, P-pg is also<br />

expressed at <strong>the</strong> luminal part <strong>of</strong> <strong>the</strong> intestinal mucosa [20,53].<br />

Its physiological function is mainly to avoid <strong>the</strong> absorption <strong>of</strong><br />

toxic xenobiotics and/or metabolites. Therefore, <strong>the</strong> induction<br />

<strong>of</strong> enterocyte P-pg may also gain clinical importance to prevent<br />

fur<strong>the</strong>r PQ absorption after oral intake, although this has still to<br />

be tested.<br />

It is well known that injury to <strong>the</strong> air–blood barrier and<br />

impairment <strong>of</strong> surfactant production in <strong>the</strong> lung can cause<br />

pulmonary edema and collapse <strong>of</strong> <strong>the</strong> fine airways. In <strong>the</strong><br />

present study PQ caused lung edema, observed by <strong>the</strong> increase<br />

<strong>of</strong> RLW, an effect that was attenuated by DEX. Histopathological<br />

analysis confirmed that animals from <strong>the</strong> PQ group<br />

revealed an interstitial edema, indicated by <strong>the</strong> existence <strong>of</strong><br />

intercellular vacuolization areas that were characterized by a<br />

minor density ultrastructure at TEM. Exuberance <strong>of</strong> interstitial<br />

edema was drastically attenuated in PQ+DEX animals (Figs.<br />

4E and 4F). In a previous study [54], <strong>the</strong> pretreatment <strong>of</strong><br />

1221<br />

animals with α-tocopherol liposomes or liposomes containing<br />

both α-tocopherol and GSH did not alter significantly <strong>the</strong> PQ<strong>induced</strong><br />

changes in RLW.<br />

In this study, it is highly probable that <strong>the</strong> anti-inflammatory<br />

effect <strong>of</strong> DEX contributed to its protective effect against PQ<strong>induced</strong><br />

lung toxicity. Indeed, according to Hybertson et al.<br />

[55], <strong>the</strong> pulmonary toxicity caused by PQ is assumed to have a<br />

connection with <strong>the</strong> activation <strong>of</strong> neutrophils. It was previously<br />

shown that <strong>the</strong> treatment <strong>of</strong> rats with DEX significantly reduced<br />

Sephadex-<strong>induced</strong> recruitment <strong>of</strong> inflammatory cells to bronchoalveolar<br />

fluid [56]. Fur<strong>the</strong>rmore, various inflammatory<br />

mediators have been found to be increased in <strong>the</strong> alveolar<br />

space during <strong>the</strong> early phase <strong>of</strong> acute respiratory distress<br />

syndrome, including tumor necrosis factor-α (TNF-α), interleukin-1β,<br />

interleukin-6, and chemokines [57]. TNF-α, a potent<br />

inflammatory mediator, triggers <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> leukotrienes<br />

and prostaglandin E2, which <strong>the</strong>n stimulate <strong>the</strong> infiltration <strong>of</strong><br />

polymorphonuclear leukocytes <strong>into</strong> <strong>the</strong> lungs. DEX has also<br />

been shown to decrease TNF-α concentrations in <strong>the</strong>


1222 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

bronchoalveolar lavage fluid <strong>of</strong> PQ-treated rats to about half<br />

those <strong>of</strong> control animals [58]. Of note, DEX improves gas<br />

exchange in PQ-treated animals by alleviation <strong>of</strong> lung damage<br />

after PQ-<strong>induced</strong> lung injury [58]. DEX presents also an<br />

inhibitory effect on ROS production by macrophages and<br />

neutrophils [59]. Because neutrophils are recruited to <strong>the</strong> lungs<br />

during <strong>the</strong> inflammatory reaction generated by PQ exposure,<br />

MPO activities were assessed. As expected, our results showed<br />

that MPO activity is markedly elevated in <strong>the</strong> lungs <strong>of</strong> animals<br />

exposed to PQ. Histopathological studies confirmed <strong>the</strong> widespread<br />

neutrophil infiltration <strong>into</strong> <strong>the</strong> lungs <strong>of</strong> <strong>the</strong>se animals.<br />

Macrophage infiltration and several NK cells were also<br />

identified in <strong>the</strong> interstitial space (Fig. 4D). DEX clearly<br />

reduced <strong>the</strong> lung infiltration by neutrophils observed in <strong>the</strong><br />

interstitial space or within capillaries neighboring endo<strong>the</strong>lial<br />

cells (Figs. 4E and 4F), an effect that could be attributed both to<br />

<strong>the</strong> anti-inflammatory and to its P-gp-inducing properties.<br />

However, taking <strong>into</strong> account that animals exposed to VER (in<br />

addition to DEX) showed a significant increase (p


develop new, specific, and more potent inducers <strong>of</strong> de novo<br />

syn<strong>the</strong>sis <strong>of</strong> P-gp. These inducers may be useful tools in <strong>the</strong><br />

reduction <strong>of</strong> systemic exposure and specific tissue access <strong>of</strong><br />

potential harmful xenobiotics, like PQ.<br />

Acknowledgment<br />

Ricardo Dinis-Oliveira acknowledges FCT for his Ph.D.<br />

grant (SFRH/BD/13707/2003).<br />

References<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

[1] Onyeama, H. P.; Oehme, F. W. A literature review <strong>of</strong> <strong>paraquat</strong> toxicity. Vet.<br />

Hum. Toxicol. 26:494–502; 1984.<br />

[2] Rose, M. S.; Smith, L. L.; Wyatt, I. Evidence for energy-dependent<br />

accumulation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong> rat lung. Nature 252:314–315; 1974.<br />

[3] Rannels, D. E.; Pegg, A. E.; Clark, R. S.; Addison, J. L. Interaction <strong>of</strong><br />

<strong>paraquat</strong> and amine uptake by rat lungs perfused in situ. Am. J. Physiol.<br />

Endocrinol. Metab. 249:E506–E513; 1985.<br />

[4] Dinis-Oliveira, R. J.; Valle, M. J. d. J.; Bastos, M. L.; Carvalho, F.;<br />

Sánchez-Navarro, A. Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung: influence<br />

<strong>of</strong> sodium depletion. Xenobiotica (in press); 2006.<br />

[5] Clejan, L.; Cederbaum, A. I. Synergistic interaction between NADPHcytochrome<br />

P-450 reductase, <strong>paraquat</strong> and iron in <strong>the</strong> generation <strong>of</strong> active<br />

oxygen radicals. Biochem. Pharmacol. 38:1779–1786; 1989.<br />

[6] Dinis-Oliveira, R. J.; Remião, F.; Duarte, J. A.; Sánchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. Paraquat exposure as an etiological factor <strong>of</strong><br />

Parkinson’s disease. Neurotoxicology (in press); 2006.<br />

[7] Fukushima, T.; Yamada, K.; Isobe, A.; Shiwaku, K.; Yamane, Y.<br />

Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. I. NADH oxidation and <strong>paraquat</strong><br />

radical formation via complex I. Exp. Toxicol. Pathol. 45:345–349;<br />

1993.<br />

[8] Yamada, K.; Fukushima, T. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. II.<br />

Organ specificity <strong>of</strong> <strong>paraquat</strong>-stimulated lipid peroxidation in <strong>the</strong> inner<br />

membrane <strong>of</strong> mitochondria. Exp. Toxicol. Pathol. 45:375–380; 1993.<br />

[9] Bus, J. S.; Aust, S. D.; Gibson, J. E. Superoxide- and singlet oxygencatalyzed<br />

lipid peroxidation as a possible mechanism for <strong>paraquat</strong> (methyl<br />

viologen) toxicity. Biochem. Biophys. Res. Commun. 58:749–755; 1974.<br />

[10] Dicker, E.; Cederbaum, A. I. NADH-dependent generation <strong>of</strong> reactive<br />

oxygen species by microsomes in <strong>the</strong> presence <strong>of</strong> iron and redox cycling<br />

agents. Biochem. Pharmacol. 42:529–535; 1991.<br />

[11] Bateman, D. N. Pharmacological treatments <strong>of</strong> <strong>paraquat</strong> poisoning. Hum.<br />

Toxicol. 6:57–62; 1987.<br />

[12] Bismuth, C.; Garnier, R.; Dally, S.; Fournier, P. E.; Scherrmann, J. M.<br />

Prognosis and treatment <strong>of</strong> <strong>paraquat</strong> poisoning: a review <strong>of</strong> 28 cases.<br />

J. Toxicol. Clin. Toxicol. 19:461–474; 1982.<br />

[13] Meredith, T. J.; Vale, J. A. Treatment <strong>of</strong> <strong>paraquat</strong> poisoning in man:<br />

methods to prevent absorption. Hum. Toxicol. 6:49–55; 1987.<br />

[14] Kohen, R.; Chevion, M. Transition metals potentiate <strong>paraquat</strong> toxicity.<br />

Free Radic. Res. Commun. 1:79–88; 1985.<br />

[15] Suntres, Z. E. Role <strong>of</strong> antioxidants in <strong>paraquat</strong> toxicity. Toxicology<br />

180:65–77; 2002.<br />

[16] Chen, G. H.; Lin, J. L.; Huang, Y. K. Combined methylprednisolone and<br />

dexamethasone <strong>the</strong>rapy for <strong>paraquat</strong> poisoning. Crit. Care Med. 30:<br />

2584–2587; 2002.<br />

[17] Lin, J. L.; Leu, M. L.; Liu, Y. C.; Chen, G. H. A prospective clinical trial <strong>of</strong><br />

pulse <strong>the</strong>rapy with glucocorticoid and cyclophosphamide in moderate to<br />

severe <strong>paraquat</strong>-poisoned patients. Am. J. Respir. Crit. Care Med. 159:<br />

357–360; 1999.<br />

[18] Buss, D. S.; McCaffery, A. R.; Callaghan, A. Evidence for p-glycoprotein<br />

modification <strong>of</strong> insecticide toxicity in mosquitoes <strong>of</strong> <strong>the</strong> Culex pipiens<br />

complex. Med. Vet. Entomol. 16:218–222; 2002.<br />

[19] Cordon-Cardo, C.; O’Brien, J.; Boccia, J.; Casals, D.; Bertino, J.; Melamed,<br />

M. Expression <strong>of</strong> <strong>the</strong> multidrug resistance gene product (P-glycoprotein) in<br />

human normal and tumor tissues. J. Histochem. Cytochem. 38:1277–1287;<br />

1990.<br />

1223<br />

[20] Gottesman, M. M.; Pastan, I. Biochemistry <strong>of</strong> multidrug resistance<br />

mediated by <strong>the</strong> multidrug transporter. Annu. Rev. Biochem. 62:385–427;<br />

1993.<br />

[21] Ruetz, S.; Gros, P. Functional expression <strong>of</strong> P-glycoproteins in secretory<br />

vesicles. J. Biol. Chem. 269:12277–12284; 1994.<br />

[22] Demeule, M.; Jodoin, J.; Beaulieu, E.; Brossard, M.; Beliveau, R.<br />

Dexamethasone modulation <strong>of</strong> multidrug transporters in normal tissues.<br />

FEBS Lett. 442:208–214; 1999.<br />

[23] Stein, W. D. Kinetics <strong>of</strong> <strong>the</strong> multidrug transporter (P-glycoprotein) and its<br />

reversal. Physiol. Rev. 77:545–590; 1997.<br />

[24] Choi, J. S.; Li, X. The effect <strong>of</strong> verapamil on <strong>the</strong> pharmacokinetics <strong>of</strong><br />

paclitaxel in rats. Eur. J. Pharm. Sci. 24:95–100; 2005.<br />

[25] Akahori, F.; Masaoka, T.; Matsushiro, S.; Arishima, K.; Arai, S.;<br />

Yamamoto, M.; Eguchi, Y. Quantifiable morphologic evaluation <strong>of</strong><br />

<strong>paraquat</strong> pulmonary toxicity in rats. Vet. Hum. Toxicol. 29:1–7; 1987.<br />

[26] Rocco, P. R.; Souza, A. B.; Faffe, D. S.; Passaro, C. P.; Santos, F. B.; Negri,<br />

E. M.; Lima, J. G.; Contador, R. S.; Capelozzi, V. L.; Zin, W. A. Effect <strong>of</strong><br />

corticosteroid on lung parenchyma remodeling at an early phase <strong>of</strong> acute<br />

lung injury. Am. J. Respir. Crit. Care Med. 168:677–684; 2003.<br />

[27] Fuke, C.; Ameno, K.; Ameno, S.; Kiriu, T.; Shinohara, T.; Sogo, K.; Ijiri, I.<br />

A rapid, simultaneous determination <strong>of</strong> <strong>paraquat</strong> and diquat in serum and<br />

urine using second-derivative spectroscopy. J. Anal. Toxicol. 16:214–216;<br />

1992.<br />

[28] Laemmli, U. K. Cleavage <strong>of</strong> structural proteins during <strong>the</strong> assembly <strong>of</strong> <strong>the</strong><br />

head <strong>of</strong> bacteriophage T4. Nature 227:680–685; 1970.<br />

[29] Duarte, J. A.; Leao, A.; Magalhaes, J.; Ascensao, A.; Bastos, M. L.;<br />

Amado, F. L.; Vilarinho, L.; Quelhas, D.; Appell, H. J.; Carvalho, F.<br />

Strenuous exercise aggravates MDMA-<strong>induced</strong> skeletal muscle damage in<br />

mice. Toxicology 206:349–358; 2005.<br />

[30] Chen, C. M.; Wang, L. F.; Su, B.; Hsu, H. H. Methylprednisolone effects<br />

on oxygenation and histology in a rat model <strong>of</strong> acute lung injury. Pulm.<br />

Pharmacol. Ther. 16:215–220; 2003.<br />

[31] Chatterjee, P. K.; Cuzzocrea, S.; Brown, P. A.; Zacharowski, K.; Stewart,<br />

K. N.; Mota-Filipe, H.; Thiemermann, C. Tempol, a membrane-permeable<br />

radical scavenger, reduces oxidant stress-mediated renal dysfunction and<br />

injury in <strong>the</strong> rat. Kidney Int. 58:658–673; 2000.<br />

[32] Ascensao, A.; Magalhaes, J.; Soares, J. M.; Ferreira, R.; Neuparth, M. J.;<br />

Marques, F.; Oliveira, P. J.; Duarte, J. A. Moderate endurance training<br />

prevents doxorubicin-<strong>induced</strong> in vivo mitochondriopathy and reduces <strong>the</strong><br />

development <strong>of</strong> cardiac apoptosis. Am. J. Physiol. Heart Circ. Physiol.<br />

289:H722–H731; 2005.<br />

[33] Lowry, O. H. N.; Rosebrough, N. J.; Farr, A. L.; Randall, R. J. Protein<br />

measurement with Folin phenol reagent. J. Biol. Chem. 193:265–275;<br />

1951.<br />

[34] Buege, J. A.; Aust, S. D. Microsomal lipid peroxidation. Methods<br />

Enzymol. 52:302–310; 1978.<br />

[35] Levine, R. L.; Williams, J. A.; Stadtman, E. R.; Shacter, E. Carbonyl assay<br />

for determination <strong>of</strong> oxidatively modified proteins. Methods Enzymol.<br />

233:346–357; 1994.<br />

[36] Andrews, P. C.; Krinsky, N. I. Quantitative determination <strong>of</strong> myeloperoxidase<br />

using tetramethylbenzidine as substrate. Anal. Biochem. 127:<br />

346–350; 1982.<br />

[37] Suzuki, K.; Ota, H.; Sasagawa, S.; Sakatani, T.; Fujikura, T. Assay method<br />

for myeloperoxidase in human polymorphonuclear leukocytes. Anal.<br />

Biochem. 132:345–352; 1983.<br />

[38] Vandeputte, C.; Guizon, I.; Genestie-Denis, I.; Vannier, B.; Lorenzon, G. A<br />

microtiter plate assay for total glutathione and glutathione disulfide<br />

contents in cultured/isolated cells: performance study <strong>of</strong> a new miniaturized<br />

protocol. Cell Biol. Toxicol. 10:415–421; 1994.<br />

[39] Flohe, L.; Otting, F. Superoxide dismutase assays. Methods Enzymol.<br />

105:93–104; 1984.<br />

[40] Michaluk, J.; Karolewicz, B.; Antkiewicz-Michaluk, L.; Vetulani, J.<br />

Effects <strong>of</strong> various Ca 2+ channel antagonists on morphine analgesia,<br />

tolerance and dependence, and on blood pressure in <strong>the</strong> rat. Eur. J.<br />

Pharmacol. 352:189–197; 1998.<br />

[41] Pratt, W. B.; T<strong>of</strong>t, D. O. Steroid receptor interactions with heat shock<br />

protein and immunophilin chaperones. Endocr. Rev. 18:306–360;<br />

1997.


1224 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 41 (2006) 1213–1224<br />

[42] Cosio, B. G.; Torrego, A.; Adcock, I. M. Molecular <strong>mechanisms</strong> <strong>of</strong><br />

glucocorticoids. Arch. Bronconeumol. 41:34–41; 2005.<br />

[43] Fardel, O.; Lecureur, V.; Guillouzo, A. Regulation by dexamethasone <strong>of</strong><br />

P-glycoprotein expression in cultured rat hepatocytes. FEBS Lett. 327:<br />

189–193; 1993.<br />

[44] Chen, C. J.; Chin, J. E.; Ueda, K.; Clark, D. P.; Pastan, I.; Gottesman,<br />

M. M.; Roninson, I. B. Internal duplication and homology with bacterial<br />

transport proteins in <strong>the</strong> mdr1 (P-glycoprotein) gene from multidrugresistant<br />

human cells. Cell 47:381–389; 1986.<br />

[45] Croop, J. M.; Raymond, M.; Haber, D.; Devault, A.; Arceci, R. J.; Gros, P.;<br />

Housman, D. E. The three mouse multidrug resistance (mdr) genes are<br />

expressed in a tissue-specific manner in normal mouse tissue. Mol. Cell.<br />

Biol. 9:1346–1350; 1989.<br />

[46] Durr, D.; Stieger, B.; Kullak-Ublick, G. A.; Rentsch, K. M.; Steinert, H. C.;<br />

Meier, P. J.; Fattinger, K. St John’s wort induces intestinal P-glycoprotein/<br />

MDR1 and intestinal and hepatic CYP3A4. Clin. Pharmacol. Ther. 68:<br />

598–604; 2000.<br />

[47] Salphati, L.; Benet, L. Z. Modulation <strong>of</strong> P-glycoprotein expression by<br />

cytochrome P450 3A inducers in male and female rat livers. Biochem.<br />

Pharmacol. 55:387–395; 1998.<br />

[48] Dinis-Oliveira, R. J.; Sarmento, A. M.; Reis, P.; Amaro, A.; Remião, F.;<br />

Bastos, M. L.; Carvalho, F. Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival<br />

case following intake <strong>of</strong> a potential lethal dose. Pediatr. Emerg. Care<br />

22:537–540; 2006.<br />

[49] Toth, G. G.; Kloosterman, C.; Uges, D. R.; Jonkman, M. F. Pharmacokinetics<br />

<strong>of</strong> high-dose oral and intravenous dexamethasone. Ther. Drug.<br />

Monit. 21:532–535; 1999.<br />

[50] Fardel, O.; Payen, L.; Courtois, A.; Vernhet, L.; Lecureur, V. Regulation <strong>of</strong><br />

biliary drug efflux pump expression by hormones and xenobiotics. Toxicology<br />

167:37–46; 2001.<br />

[51] Fardel, O.; Payen, L.; Sparfel, L.; Vernhet, L.; Lecureur, V. Drug<br />

membrane transporters in <strong>the</strong> liver: regulation <strong>of</strong> <strong>the</strong>ir expression and<br />

activity. Ann. Pharm. Fr. 60:380–385; 2002.<br />

[52] Hughes, R. D.; Millburn, P.; Williams, R. T. Biliary excretion <strong>of</strong> some<br />

diquaternary ammonium cations in <strong>the</strong> rat, guinea pig and rabbit. Biochem.<br />

J. 136:979–984; 1973.<br />

[53] Terao, T.; Hisanaga, E.; Sai, Y.; Tamai, I.; Tsuji, A. Active secretion <strong>of</strong> drugs<br />

from <strong>the</strong> small intestinal epi<strong>the</strong>lium in rats by P-glycoprotein functioning as<br />

an absorption barrier. J. Pharm. Pharmacol. 48: 1083–1089; 1996.<br />

[54] Suntres, Z. E.; Shek, P. N. Alleviation <strong>of</strong> <strong>paraquat</strong>-<strong>induced</strong> lung injury by<br />

pretreatment with bifunctional liposomes containing alpha-tocopherol and<br />

glutathione. Biochem. Pharmacol. 52:1515–1520; 1996.<br />

[55] Hybertson, B. M.; Lampey, A. S.; Clarke, J. H.; Koh, Y.; Repine, J. E. Nacetylcysteine<br />

pretreatment attenuates <strong>paraquat</strong>-<strong>induced</strong> lung leak in rats.<br />

Redox Rep. 1:337–342; 1995.<br />

[56] Williams, C. M.; Smith, L.; Flanagan, B. F.; Clegg, L. S.; Coleman, J. W.<br />

Tumour necrosis factor-alpha expression and cell recruitment in Sephadex<br />

particle-<strong>induced</strong> lung inflammation: effects <strong>of</strong> dexamethasone and cyclosporin<br />

A. Br. J. Pharmacol. 122:1127–1134; 1997.<br />

[57] Pugin, J.; Verghese, G.; Widmer, M. C.; Matthay, M. A. The alveolar<br />

space is <strong>the</strong> site <strong>of</strong> intense inflammatory and pr<strong>of</strong>ibrotic reactions in <strong>the</strong><br />

early phase <strong>of</strong> acute respiratory distress syndrome. Crit. Care Med. 27:<br />

304–312; 1999.<br />

[58] Chen, C. M.; Fang, C. L.; Chang, C. H. Surfactant and corticosteroid<br />

effects on lung function in a rat model <strong>of</strong> acute lung injury. Crit. Care Med.<br />

29:2169–2175; 2001.<br />

[59] Maridonneau-Parini, I.; Errasfa, M.; Russo-Marie, F. Inhibition <strong>of</strong> O 2 −<br />

generation by dexamethasone is mimicked by lipocortin I in alveolar<br />

macrophages. J. Clin. Invest. 83:1936–1940; 1989.<br />

[60] Fukushima, T.; Yamada, K.; Hojo, N.; Isobe, A.; Shiwaku, K.; Yamane, Y.<br />

Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>: III. The effects <strong>of</strong> acute <strong>paraquat</strong><br />

exposure on <strong>the</strong> electron transport system in rat mitochondria. Exp.<br />

Toxicol. Pathol. 46:437–441; 1994.<br />

[61] Bus, J. S.; Aust, S. D.; Gibson, J. E. Lipid peroxidation: a possible<br />

mechanism for <strong>paraquat</strong> toxicity. Res. Commun. Chem. Pathol. Pharmacol.<br />

11:31–38; 1975.<br />

[62] Burk, R. F.; Lawrence, R. A.; Lane, J. M. Liver necrosis and lipid<br />

peroxidation in <strong>the</strong> rat as result <strong>of</strong> <strong>paraquat</strong> and diquat administration:<br />

effect <strong>of</strong> selenium deficiency. J. Clin. Invest. 65:1024–1031;<br />

1980.<br />

[63] Dean, R. T.; Fu, S.; Stocker, R.; Davies, M. J. Biochemistry and<br />

pathology <strong>of</strong> radical-mediated protein oxidation. Biochem. J. 324:1–18;<br />

1997.<br />

[64] Seelig, G. F.; Meister, A. Gamma-glutamylcysteine syn<strong>the</strong>tase: interactions<br />

<strong>of</strong> an essential sulfhydryl group. J. Biol. Chem. 259:3534–3538;<br />

1984.<br />

[65] Cantin, A. M.; North, S. L.; Hubbard, R. C.; Crystal, R. G. Normal alveolar<br />

epi<strong>the</strong>lial lining fluid contains high levels <strong>of</strong> glutathione. J. Appl. Physiol.<br />

63:152–157; 1987.<br />

[66] Li, X. Y.; Donaldson, K.; Rahman, I.; MacNee, W. An investigation <strong>of</strong> <strong>the</strong><br />

role <strong>of</strong> glutathione in increased epi<strong>the</strong>lial permeability <strong>induced</strong> by cigarette<br />

smoke in vivo and in vitro. Am. J. Respir. Crit. Care Med. 149:<br />

1518–1525; 1994.<br />

[67] Maellaro, E.; Casini, A. F.; Bello, B. D.; Comporti, M. Lipid peroxidation<br />

and antioxidant systems in <strong>the</strong> liver injury produced by glutathione<br />

depleting agents. Biochem. Pharmacol. 39:1513–1521; 1990.<br />

[68] Keeling, P. L.; Smith, L. L. Relevance <strong>of</strong> NADPH depletion and mixed<br />

disulphide formation in rat lung to <strong>the</strong> mechanism <strong>of</strong> cell damage<br />

following <strong>paraquat</strong> administration. Biochem. Pharmacol. 31:3243–3249;<br />

1982.


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER IV<br />

Single high dose dexamethasone treatment decreases <strong>the</strong> pathological effects and<br />

increases <strong>the</strong> survival rat <strong>of</strong> <strong>paraquat</strong>-<strong>into</strong>xicated rats<br />

Reprinted from Toxicology 227: 73-85<br />

Copyright© (2006) with kind permission from Elsevier Science Inc<br />

161


Part II – Original <strong>research</strong>____________________________________________________<br />

162


Abstract<br />

Toxicology 227 (2006) 73–85<br />

Single high dose dexamethasone treatment decreases <strong>the</strong><br />

pathological score and increases <strong>the</strong> survival rate <strong>of</strong><br />

<strong>paraquat</strong>-<strong>into</strong>xicated rats<br />

R.J. Dinis-Oliveira a,∗ , J.A. Duarte b , F. Remião a ,A.Sánchez-Navarro c ,<br />

M.L. Bastos a ,Félix Carvalho a,∗<br />

a REQUIMTE, Departamento de Toxicologia, Faculdade de Farmácia, Universidade do Porto,<br />

Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal<br />

b Departamento de Biologia do Desporto, Faculdade de Ciências do Desporto, Universidade do Porto,<br />

Rua Dr. Plácido Costa, 91, 4200-450 Porto, Portugal<br />

c Departamento de Farmacia y Tecnología Farmacéutica, Facultad de Farmacia, Universidad de Salamanca,<br />

Avda. Campo Charro s/n, 37007, Salamanca, Spain<br />

Received 20 June 2006; received in revised form 13 July 2006; accepted 14 July 2006<br />

Available online 3 August 2006<br />

Dexamethasone (DEX), a syn<strong>the</strong>tic corticosteroid, has been successfully used in clinical practice during <strong>paraquat</strong> (PQ) poisonings<br />

due to its anti-inflammatory activity, although, as recently observed, its effects related to de novo syn<strong>the</strong>sis <strong>of</strong> P-glycoprotein (P-gp),<br />

may also strongly contribute for its healing effects. The main purpose <strong>of</strong> this study was to evaluate <strong>the</strong> effects <strong>of</strong> a single high dose<br />

DEX administration, which induces de novo syn<strong>the</strong>sis <strong>of</strong> P-gp, in <strong>the</strong> histological and biochemical parameters in lung, liver, kidney<br />

and spleen <strong>of</strong> acute PQ-<strong>into</strong>xicated rats. Four groups <strong>of</strong> rats were constituted: (i) control group, (ii) DEX group (100 mg/kg i.p.), (iii)<br />

PQ group (25 mg/kg i.p.) and (iv) PQ + DEX group (DEX injected 2 h after PQ). The obtained results showed that DEX ameliorated<br />

<strong>the</strong> biochemical and histological lung and liver alterations <strong>induced</strong> by PQ in Wistar rats at <strong>the</strong> end <strong>of</strong> 24 hours. This was evidenced by<br />

a significant reduction in lipid peroxidation (LPO) and carbonyl groups content, as well as by normalization <strong>of</strong> <strong>the</strong> myeloperoxidase<br />

(MPO) activities. Moreover, DEX prevented <strong>the</strong> increase <strong>of</strong> relative lung weight. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong>se improvements were not<br />

observed in kidney and spleen <strong>of</strong> DEX treated rats. Conversely, an increase <strong>of</strong> LPO and carbonyl groups content and aggravation<br />

<strong>of</strong> histological damages were observed in <strong>the</strong> latter tissues. In addition, MPO activity increased in <strong>the</strong> spleen <strong>of</strong> PQ + DEX group<br />

and urinary N-acetyl-�-d-glucosaminidase activity, a biomarker <strong>of</strong> renal tubular proximal damage, also augmented in this group.<br />

Never<strong>the</strong>less, it is legitimate to hypo<strong>the</strong>size that <strong>the</strong> apparent protection <strong>of</strong> high dosage DEX treatment awards to <strong>the</strong> lungs <strong>of</strong> <strong>the</strong><br />

PQ-<strong>into</strong>xicated animals outweighs <strong>the</strong> increased damage to <strong>the</strong>ir spleens and kidneys, because a higher survival rate was observed,<br />

indicating that DEX treatment may constitute an important and valuable <strong>the</strong>rapeutic drug to be used against PQ-<strong>induced</strong> toxicity.<br />

© 2006 Elsevier Ireland Ltd. All rights reserved.<br />

Keywords: Paraquat; Dexamethasone; Oxidative damage; Rats; Lung; Kidney; Liver; Spleen<br />

Abbreviations: DEX, dexamethasone; DNPH, 2,4-dinitrophenylhydrazine; H2O2, hydrogen peroxide; LM, light microscopy; LPO, lipid<br />

peroxidation; MDA, malondialdehyde; MPO, myeloperoxidase; NAG, N-acetyl-�-d-glucosaminidase; P-gp, P-glycoprotein; PALS, periarteriolar<br />

lymphocyte sheath; PQ, <strong>paraquat</strong>; RKW, relative kidney weight; RLW, relative lung weight; RLiW, relative liver weight; ROS, reactive oxygen<br />

species; ROW, relative <strong>organ</strong> weight; RSW, relative spleen weight; TBARS, thiobarbituric acid reactive substances; TCA, trichloroacetic acid; TEM,<br />

transmission electron microscopy; TMB, 3,3 ′ ,5,5 ′ -tetramethylbenzidine; TNF-�, tumor necrosis factor-alpha; VER, verapamil<br />

∗ Corresponding authors. Tel.: +351 222078922; fax: +351 222003977.<br />

E-mail addresses: ricardinis@ff.up.pt (R.J. Dinis-Oliveira), felixdc@ff.up.pt (F. Carvalho).<br />

0300-483X/$ – see front matter © 2006 Elsevier Ireland Ltd. All rights reserved.<br />

doi:10.1016/j.tox.2006.07.025


74 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

1. Introduction<br />

Since its introduction in agriculture in 1962, <strong>the</strong><br />

widespread non-selective contact herbicide <strong>paraquat</strong><br />

(PQ) used as desiccant and defoliant in a variety <strong>of</strong> crops<br />

has caused thousands <strong>of</strong> deaths in humans from both<br />

accidental and voluntary exposure. It may be considered<br />

one <strong>of</strong> <strong>the</strong> most toxic poisons involved in suicide<br />

attempts. Never<strong>the</strong>less, it is readily available without<br />

legal restrictions in several countries where it is registered,<br />

due to its herbicide effectiveness and its rapid<br />

inactivation in <strong>the</strong> environment.<br />

Since antidotes for PQ are unknown, over <strong>the</strong> past<br />

60 years strategies in <strong>the</strong> management <strong>of</strong> PQ poisonings<br />

have been directed towards modification <strong>of</strong> its<br />

toxicokinetics ei<strong>the</strong>r by decreasing <strong>the</strong> absorption or by<br />

enhancing its elimination (Dinis-Oliveira et al., 2006a).<br />

Besides <strong>the</strong>se approaches, additional protective protocols<br />

have also been adopted, particularly those aimed to<br />

reduce inflammation (Chen et al., 2002). Indeed, it has<br />

been proven that <strong>the</strong> anti-inflammatory corticosteroid<br />

<strong>the</strong>rapy reduces morbidity and mortality if used at an<br />

early phase <strong>of</strong> PQ-<strong>induced</strong> acute lung injury by ameliorating<br />

<strong>the</strong> respiratory <strong>mechanisms</strong>, lung histology and<br />

<strong>the</strong> structural remodelling <strong>of</strong> lung parenchyma in rats<br />

(Rocco et al., 2003). Dexamethasone [DEX (a syn<strong>the</strong>tic<br />

glucocorticoid)] has been successfully used in <strong>the</strong> clinical<br />

treatment <strong>of</strong> PQ poisonings (Chen et al., 2002; Dinis-<br />

Oliveira et al., 2006a), its positive effects being attributed<br />

to <strong>the</strong> down-regulation <strong>of</strong> neutrophils recruitment, collagenase<br />

activity and proliferation <strong>of</strong> type II pneumocytes<br />

(Meduri et al., 1991). Recently, our group demonstrated<br />

that <strong>the</strong> protection afforded with DEX could also be<br />

explained by <strong>the</strong> overexpression <strong>of</strong> P-glycoprotein<br />

(P-gp) in <strong>the</strong> cytoplasmic membrane, leading to <strong>the</strong><br />

elimination <strong>of</strong> PQ from lung cells and subsequent faecal<br />

excretion (Dinis-Oliveira et al., 2006b). Currently, <strong>the</strong>se<br />

clinical beneficial effects are mainly supported by <strong>the</strong><br />

subjacent DEX protective <strong>mechanisms</strong> described in<br />

lungs. However, beyond lung, PQ accumulation has<br />

also been observed in o<strong>the</strong>r <strong>organ</strong>s such as kidney, liver<br />

and spleen (Sharp et al., 1972). In addition, histological<br />

and biochemical modifications, suggestive <strong>of</strong> oxidative<br />

stress and damage, have also been described in such<br />

<strong>organ</strong>s after acute PQ exposure (Akahori et al., 1987;<br />

Burk et al., 1980; Lock and Ishmael, 1979; Melchiorri<br />

et al., 1996). In fact, with high ingestion doses <strong>of</strong> PQ<br />

(>30 mg/kg in humans), death occurs within 1 week<br />

after <strong>into</strong>xication resulting from <strong>multiple</strong> <strong>organ</strong> failure<br />

(Bismuth et al., 1990; Onyeama and Oehme, 1984).<br />

Considering that <strong>the</strong> extrapulmonary repercussions<br />

<strong>of</strong> DEX <strong>the</strong>rapy in PQ <strong>into</strong>xications, relatively to its<br />

biochemical and histological effects, still remain poorly<br />

understood, <strong>the</strong> aim <strong>of</strong> this work was to provide comprehensive<br />

results about <strong>the</strong> effect <strong>of</strong> DEX administration<br />

on inflammatory reaction, oxidative stress and<br />

related damage, assessed by histological and biochemical<br />

parameters in lung, liver, kidney and spleen <strong>of</strong> acute<br />

PQ-<strong>into</strong>xicated rats. Moreover, it was also our objective<br />

to evaluate <strong>the</strong> overall healing provided by DEX as well<br />

as <strong>the</strong> hypo<strong>the</strong>tic contribution <strong>of</strong> P-gp de novo syn<strong>the</strong>sis<br />

to that protection by presenting <strong>the</strong> survival rate curves.<br />

2. Materials and methods<br />

2.1. Chemicals and drugs<br />

Paraquat dichloride (1,1 ′ -dimethyl-4,4 ′ -bipyridinium dichloride),<br />

dexamethasone [(11�,16�)-9-fluoro-11,17,21-trihydroxy-16-methylpregna-1,4-diene-3,20-dione],3,3’,5,5’tetramethylbenzidine<br />

(TMB), 4-nitrophenyl N-acetyl-�-dglucosaminide,<br />

2-amino-2-methyl-1-propanol hydrochloride<br />

and 2,4-dinitrophenylhydrazine (DNPH) were all obtained<br />

from Sigma (St. Louis, MO, U.S.A.). The saline solution<br />

(NaCl 0.9%), sodium thiopental were obtained from B. Braun<br />

(Lisbon, Portugal). 2-Thiobarbituric acid (C4H4N2O2S),<br />

trichloroacetic acid (TCA; Cl3CCOOH) and sodium hydroxide<br />

(NaOH) were obtained from Merck (Darmstadt, Germany).<br />

All <strong>the</strong> reagents used were <strong>of</strong> analytical grade or from <strong>the</strong><br />

highest available grade.<br />

2.2. Animals<br />

The study was performed in two steps, both using adult<br />

male Wistar rats (aged 8 weeks) obtained from Charles<br />

River S.A. (Barcelona, Spain), with a mean body weight <strong>of</strong><br />

252 ± 10 g. Animals were kept in standard laboratory conditions<br />

(12/12 h light/darkness, 22 ± 2 ◦ C room temperature,<br />

50–60% humidity) for at least 1 week before starting <strong>the</strong><br />

experiments. Animals were allowed access to tap water and<br />

rat chow ad libitum during <strong>the</strong> quarantine period. Animal<br />

experiments were licensed by Portuguese General Directorate<br />

<strong>of</strong> Veterinary Medicine. Housing and experimental treatment<br />

<strong>of</strong> animals were in accordance with <strong>the</strong> Guide for <strong>the</strong> Care and<br />

Use <strong>of</strong> Laboratory Animals from <strong>the</strong> Institute for Laboratory<br />

Animal Research (ILAR, 1996). The experiments complied<br />

with <strong>the</strong> current Portuguese laws.<br />

2.3. Experimental protocol for biochemical and<br />

histological studies<br />

The biochemical and histological studies were carried out<br />

in 32 animals randomly divided <strong>into</strong> four groups. Each animal<br />

was individually housed in a metabolic cage and kept during<br />

<strong>the</strong> experiment (26 h) for whole urine collection. Animals<br />

were fasted during <strong>the</strong> entire experimental period but water<br />

was given ad libitum.


The four groups were treated as follows (given doses<br />

were kg per body weight): (i) control group, n = 8: animals<br />

treated with 0.9% NaCl. Animals received one more administration<br />

<strong>of</strong> 0.9% NaCl 2 h later. (ii) DEX group, n = 8: animals<br />

treated with DEX (100 mg/kg). Animals received one<br />

administration <strong>of</strong> 0.9% NaCl 2 h later (iii) PQ group, n =8:<br />

animals <strong>into</strong>xicated with PQ (25 mg/kg). Animals received<br />

one administration <strong>of</strong> 0.9% NaCl 2 h later. (iv) PQ + DEX<br />

group, n = 8: animals <strong>into</strong>xicated with PQ (25 mg/kg). Two<br />

hours later, animals were treated with DEX (100 mg/kg). The<br />

schedule <strong>of</strong> DEX administration was chosen considering <strong>the</strong><br />

arrival time <strong>of</strong> <strong>the</strong> patient to <strong>the</strong> hospital, after PQ <strong>into</strong>xication.<br />

The administrations <strong>of</strong> vehicle (0.9% NaCl), PQ and<br />

DEX were all made intraperitoneally (i.p.) in an injection<br />

volume <strong>of</strong> 0.5 ml. PQ dose was similar to that used in previous<br />

studies, conducting to severe lung toxicity (Akahori et<br />

al., 1987; Rocco et al., 2003). The reported LD50 for rats in<br />

<strong>the</strong> literature is ∼18–28 mg/kg <strong>of</strong> PQ dichloride (Clark et al.,<br />

1966).<br />

Treatments in all groups were always conducted between<br />

8:00 and 10:00 a.m.<br />

2.4. Surgical procedures<br />

Twenty-six hours after <strong>the</strong> first injection, anes<strong>the</strong>sia was<br />

<strong>induced</strong> with sodium thiopental (60 mg/kg, i.p.). Animals were<br />

placed in <strong>the</strong> decubito supino position and abdomen was<br />

opened by two lateral transversal incisions and one central<br />

longitudinal incision to expose <strong>the</strong> portal vein. Five animals<br />

<strong>of</strong> each group (biochemical determinations) were perfused in<br />

situ with ice-cold 0.9% NaCl for 3 min at a rate <strong>of</strong> 10 ml/min<br />

through <strong>the</strong> portal vein and completely cleaned <strong>of</strong> blood. In <strong>the</strong><br />

remaining three animals (histological analysis), <strong>organ</strong>s’ perfusion<br />

was done with 2.5% glutaraldehyde in 0.2 M sodium<br />

cacodylate buffer (pH 7.2–7.4) in order to pre-fixate tissues for<br />

fur<strong>the</strong>r histological analysis. Simultaneously to <strong>the</strong> perfusion<br />

initiation, a cut at <strong>the</strong> common iliac arteries was done to avoid<br />

overpressure.<br />

2.5. Collection and processing samples for biochemical<br />

measurements<br />

Organs were removed, pat-dried with gauze, weighted and<br />

processed as following: (i) right lung, right kidney, half-spleen<br />

and liver right lobe were homogenized (1:4, m/v, Ultra-Turrax ®<br />

Homogenizer) in ice-cold 50 mM phosphate buffer with 0.1%<br />

(v/v) Triton X-100 and at pH 7.4. The homogenate was kept<br />

on ice, <strong>the</strong>n centrifuged at 3000 × g,4 ◦ C, for 10 min. Aliquots<br />

<strong>of</strong> <strong>the</strong> resulting supernatant’s were stored (−80 ◦ C) for posterior<br />

quantification <strong>of</strong> myeloperoxidase activity (MPO), carbonyl<br />

groups, PQ and protein content. (ii) The left lung,<br />

left kidney, half-spleen and liver left lobe were homogenized<br />

(1:4 m/v, Ultra-Turrax ® Homogenizer) in TCA 10%. The<br />

homogenate was kept on ice and <strong>the</strong>n centrifuged at 13,000 × g,<br />

4 ◦ C, for 20 min. Aliquots <strong>of</strong> <strong>the</strong> resulting supernatants were<br />

immediately used for evaluating <strong>the</strong> lipid peroxidation (LPO)<br />

degree.<br />

R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 75<br />

The relative <strong>organ</strong> weight (ROW) <strong>of</strong> each animal was also<br />

calculated as a percentage <strong>of</strong> <strong>the</strong> absolute body weight at <strong>the</strong><br />

sacrifice day.<br />

2.6. Biochemical assays<br />

Protein quantification was performed according to <strong>the</strong><br />

method <strong>of</strong> Lowry et al. (1951), using bovine serum albumin<br />

as standard.<br />

LPO was evaluated by <strong>the</strong> thiobarbituric acid reactive<br />

substances (TBARS) methodology (Buege and Aust, 1978).<br />

Results were expressed as nmol <strong>of</strong> malondialdehyde (MDA)<br />

equivalents/mg protein using an extinction coefficient (ε) <strong>of</strong><br />

1.56 × 10 5 M −1 cm −1 .<br />

Carbonyl groups (ketones and aldehydes) were determined<br />

according to Levine et al. (1994). Results were expressed<br />

as nanomole <strong>of</strong> DNPH incorporated per mg <strong>of</strong> protein<br />

(ε = 2.2 × 10 4 M −1 cm −1 ).<br />

MPO activity was measured according to <strong>the</strong> method followed<br />

by Suzuki et al. (1983) and Andrews and Krinsky (1982),<br />

with slight modifications. Briefly, <strong>the</strong> supernatants were initially<br />

submitted to three cycles <strong>of</strong> snap freezing. The assay<br />

mixture consisted <strong>of</strong> 50 �l <strong>of</strong> supernatant and 50 �l <strong>of</strong>TMB<br />

(final concentration 7.5 mM) dissolved in dimethyl sulfoxide.<br />

The enzymatic activity was initiated by adding 50 �l <strong>of</strong> hydrogen<br />

peroxide [H2O2 (final concentration 1.5 mM)] dissolved<br />

in phosphate buffer (Na2HPO4·2H2O 50 mM, pH 5.4). The<br />

rate <strong>of</strong> MPO/H2O2-catalyzed oxidation <strong>of</strong> TMB was followed<br />

by recording <strong>the</strong> absorbance increase at 655 nm at 37 ◦ C during<br />

3 min. One enzyme Unit (U) was defined as <strong>the</strong> amount<br />

<strong>of</strong> enzyme capable to reduce 1 �l <strong>of</strong>H2O2/min under assay<br />

conditions. Results were expressed in enzyme U/g <strong>of</strong> protein<br />

(ε = 3.9 × 10 4 M −1 cm −1 ).<br />

Urinary N-acetyl-�-d-glucosaminidase (NAG) activity was<br />

assayed as previously reported (Carvalho et al., 1999), using <strong>the</strong><br />

molar extinction coefficient <strong>of</strong> 18.5 × 10 3 M −1 cm −1 . One Unit<br />

<strong>of</strong> NAG was defined as <strong>the</strong> amount <strong>of</strong> enzyme that releases one<br />

�mol <strong>of</strong> p-nitrophenol in <strong>the</strong> assay conditions. Results were<br />

expressed in U kg −1 day −1 .<br />

2.7. Quantification <strong>of</strong> <strong>paraquat</strong> in rat kidney, spleen and<br />

liver<br />

Aliquots <strong>of</strong> <strong>the</strong> right lung and kidney, half-spleen and liver<br />

right lobe supernatants were treated with 5-sulfosalicylic acid<br />

(5% in final volume) and <strong>the</strong>n centrifuged (13,000 × g,4 ◦ C for<br />

10 min). The resulting supernatant fractions were alkalinized<br />

with NaOH 10N (pH >9) and <strong>the</strong>n gently mixed with few crystals<br />

<strong>of</strong> a reductant (sodium dithionite) to give <strong>the</strong> blue color,<br />

characteristic <strong>of</strong> <strong>the</strong> PQ cation radical. PQ quantification was<br />

carried out by a previously reported method based on secondderivative<br />

spectrophotometry (Fuke et al., 1992).<br />

2.8. Tissue processing for histological analysis<br />

After <strong>the</strong> in situ prefixation, lungs, liver, kidneys and spleen<br />

were removed, sectioned <strong>into</strong> ∼1mm 3 cubic pieces, and sub-


76 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

jected to routine procedures for light microscopy (LM) and<br />

transmission electron microscopy (TEM) analysis. Fixation<br />

was continued (by diffusion) in <strong>the</strong> same fixative for 2 h.<br />

After two washing steps, <strong>of</strong> 30 min each, with buffer solution,<br />

<strong>the</strong> specimens were dehydrated in graded alcohol for 2 h,<br />

and <strong>the</strong>n embedded in Epon. Propylene oxide was <strong>the</strong> compound<br />

used in <strong>the</strong> dehydratation-impregnation transition. The<br />

inclusion phase lasted 2 days. All <strong>the</strong> procedures were carried<br />

out at 4 ◦ C, with exception <strong>of</strong> <strong>the</strong> inclusion phase, which was<br />

performed at 60 ◦ C. Subsequent to <strong>the</strong> resin polymerization,<br />

semi-thin sections (thickening 1 �m) and ultra-thin sections<br />

(500 ˚A <strong>of</strong> thickness) were prepared (Ultracut, Leica), respectively,<br />

for LM and TEM analysis. The grids, mounted with<br />

<strong>the</strong> ultra-thin specimens sections, were double-contrasted with<br />

0.5% saturated uranyl acetate aqueous solution during 30 min<br />

and <strong>the</strong>n with 0.2% lead citrate solution for 15 min. The slides,<br />

mounted with semi-thin sections, were stained with toluidine<br />

blue. Five slides and three grids from each animal (standing<br />

15 slides and 9 grids per group), were examined in a Zeiss<br />

Phomi III photomicroscope and in a transmission electronic<br />

microscope (Zeiss EM 10A).<br />

Histopathological evidences <strong>of</strong> acute tissue damage were<br />

semi-quantified according to <strong>the</strong> methodology described elsewhere<br />

(Ascensao et al., 2005; Chatterjee et al., 2000; Chen et<br />

al., 2003; Duarte et al., 2005). For each group, at least more<br />

than 1000 cells per slide and 100 cells per grid were analyzed in<br />

a blind fashion in order to semi-quantify <strong>the</strong> severity and incidence<br />

<strong>of</strong> <strong>the</strong> following parameters in every slide or grid: (i) cellular<br />

degeneration, (ii) interstitial inflammatory cell infiltration,<br />

(iii) necrotic zones and (iv) tissue dis<strong>organ</strong>ization. Considering<br />

<strong>the</strong> cellular degeneration, its severity was scored according to<br />

<strong>the</strong> number <strong>of</strong> cells showing any alterations (dilatation, vacuolization,<br />

pyknotic nuclei and cellular density) in <strong>the</strong> LM<br />

visual field: grade 0 = no change from normal; grade 1 = a limited<br />

number <strong>of</strong> isolated cells (until 5% <strong>of</strong> <strong>the</strong> total cell number);<br />

grade 2 = groups <strong>of</strong> cells (5–30% <strong>of</strong> <strong>the</strong> total cell number) and<br />

grade 3 = diffuse cell damage (higher than 30% <strong>of</strong> <strong>the</strong> total cell<br />

number). The severity <strong>of</strong> inflammatory reaction was scored<br />

<strong>into</strong>: grade 0 = no cellular infiltration; grade 1 = mild leukocyte<br />

infiltration (1–3 cells by visual field); grade 2 = moderate infiltration<br />

(4–6 leukocytes by visual field) and grade 3 = heavy<br />

infiltration by neutrophils. The severity <strong>of</strong> necrosis was scored<br />

as follows: grade 0 = no necrosis; grade 1 = dispersed necrotic<br />

foci; grade 2 = confluence necrotic areas and grade 3 = massive<br />

necrosis. The severity <strong>of</strong> tissue dis<strong>organ</strong>ization was scored<br />

according to <strong>the</strong> percentage <strong>of</strong> <strong>the</strong> affected tissue: grade<br />

0 = normal structure; grade 1 = less than one third <strong>of</strong> tissue;<br />

grade 2 = greater than one third and less than two-thirds and<br />

grade 3 = greater <strong>of</strong> two-thirds <strong>of</strong> tissue. For each animal, <strong>the</strong><br />

highest possible tissue score was 12 and <strong>the</strong> lowest was 0.<br />

2.9. Experimental protocol for <strong>the</strong> evaluation <strong>of</strong> survival<br />

rate<br />

For <strong>the</strong> evaluation <strong>of</strong> survival rate, 24 animals were randomly<br />

divided <strong>into</strong> four groups <strong>of</strong> 6 animals each. It was<br />

established a control group, a PQ group, a PQ + DEX group<br />

and a PQ + VER + DEX group. In this last group, verapamil<br />

(VER, a P-gp inhibitor) was included in attempt to assess <strong>the</strong><br />

DEX contributory effect by induction P-gp de novo syn<strong>the</strong>sis.<br />

Animals were kept in a number <strong>of</strong> three per polypropylene<br />

cage with a stainless steel net at <strong>the</strong> top and wood chips at <strong>the</strong><br />

screen bottom. Tap water and rat chow were given ad libitum<br />

during <strong>the</strong> entire experiment. The control, PQ and PQ + DEX<br />

groups were treated as described for biochemical and histological<br />

studies, with a slight modification: 1 h after <strong>the</strong> first<br />

injection <strong>the</strong> animals received an additional 0.9% NaCl i.p.<br />

administration. A fourth group, receiving verapamil (VER)<br />

was included for this experimental protocol. We have previously<br />

demonstrated (Dinis-Oliveira et al., 2006b) that <strong>the</strong><br />

induction <strong>of</strong> <strong>the</strong> P-gp de novo syn<strong>the</strong>sis by DEX decreases PQ<br />

lung accumulation and consequently its toxicity and also that<br />

VER, a competitive inhibitor <strong>of</strong> this transporter blocked DEX<br />

protective effects, causing instead an increase <strong>of</strong> PQ lung concentration<br />

and an aggravation <strong>of</strong> toxicity. Therefore, to assess<br />

<strong>the</strong> DEX contributory effect by inducing P-gp de novo syn<strong>the</strong>sis<br />

and thus study <strong>the</strong> importance <strong>of</strong> P-gp in <strong>the</strong> PQ mortality rate,<br />

animals <strong>of</strong> <strong>the</strong> fourth group (PQ + DEX + VER) were <strong>into</strong>xicated<br />

with PQ (25 mg/kg) and treated with VER (10 mg/kg)<br />

and DEX (100 mg/kg), 1 and 2 h later (i.p., 0.5 ml <strong>of</strong> 0.9%<br />

NaCl), respectively. The survival rate was registered every day<br />

until <strong>the</strong> 10th day.<br />

2.10. Statistical analysis<br />

Results are expressed as mean ± standard error <strong>of</strong> <strong>the</strong> mean<br />

(S.E.M.). Statistical comparison between groups was estimated<br />

using <strong>the</strong> non-parametric method <strong>of</strong> Kruskal–Wallis followed<br />

by <strong>the</strong> Dunn’s test. Comparison <strong>of</strong> <strong>the</strong> survival curves was<br />

performed using <strong>the</strong> Logrank test. In all cases, p-values lower<br />

than 0.05 were considered as statistically significant.<br />

3. Results<br />

The exposure <strong>of</strong> rats to PQ resulted in histological and<br />

biochemical changes in liver, kidneys, spleen and lungs.<br />

3.1. Structural and ultrastructural analysis<br />

Lung—major qualitative structural and ultrastructural<br />

alterations are depicted in Fig. 1. The respective<br />

semi-quantitative analysis is shown in Table 1. Animals<br />

from control and DEX groups presented a normal pulmonary<br />

structure at LM and TEM, without evidences<br />

<strong>of</strong> alveolar collapse or cellular infiltrations. PQ administration<br />

<strong>induced</strong> marked alterations compared to <strong>the</strong><br />

control pattern, mainly characterized by a diffuse alveoli<br />

collapse with an increased thickness <strong>of</strong> its walls.<br />

An intense vascular congestion with numerous activated<br />

platelets (suggested by changes <strong>of</strong> discoid shape,


R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 77<br />

Fig. 1. Optical (above) and electron (below) micrographs from lungs <strong>of</strong> control (A and E), dexamethasone (B and F), <strong>paraquat</strong> (C and G) and <strong>paraquat</strong><br />

plus dexamethasone (D and H) groups. A, B, E and F evidenced a normal structure and ultrastructure with <strong>the</strong> presence <strong>of</strong> some pneumocytes type<br />

II; C and G depict <strong>the</strong> alveolar collapse (*) with signs <strong>of</strong> interstitial edema (blue arrows); several infiltrative macrophages (red arrows) and<br />

polymorphonuclear cells (green arrows) adherent to endo<strong>the</strong>lium can also be observed; <strong>the</strong> alveolar collapse, cellular debris (#) and macrophages in<br />

<strong>the</strong> alveolar space is present in D; In H is depicted a necrotic cell and cellular debris within <strong>the</strong> alveolar space. (For interpretation <strong>of</strong> <strong>the</strong> references<br />

to color in this figure legend, <strong>the</strong> reader is referred to <strong>the</strong> web version <strong>of</strong> this article.)<br />

Table 1<br />

Semi-quantitative analysis <strong>of</strong> <strong>the</strong> morphological injury parameters <strong>of</strong> <strong>the</strong> control, dexamethasone (DEX), <strong>paraquat</strong> (PQ) and <strong>paraquat</strong> plus dexamethasone<br />

(PQ + DEX) groups<br />

Organ Group Evaluated morphological parameter<br />

Cell degeneration Interstitial inflammatory<br />

cell infiltration<br />

Necrotic zones Tissue dis<strong>organ</strong>ization<br />

Lung Control 0.06 ± 0.06 0.0 ± 0.0 0.0 ± 0.0 0.05 ± 0.05<br />

DEX 0.09 ± 0.06 0.0 ± 0.0 0.0 ± 0.0 0.1 ± 0.07<br />

PQ 2.00 ± 0.15 a 2.05 ± 0.18 a 1.20 ± 0.09 a 2.14 ± 0.20 a<br />

PQ + DEX 1.26 ± 0.10 b,c 0.11 ± 0.08 b,c 1.00 ± 0.17 b 1.2 ± 0.11 b,c<br />

Liver Control 0.17 ± 0.08 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0<br />

DEX 0.12 ± 0.06 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0<br />

PQ 1.30 ± 0.12 a 0.45 ± 0.11 a 1.40 ± 0.15 a 1.00 ± 0.15 a<br />

PQ + DEX 1.26 ± 0.10 b 0.11 ± 0.08 b,c 1.00 ± 0.17 b,c 1.2 ± 0.11 b<br />

Spleen Control 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0<br />

DEX 0.82 ± 0.18 a 0.0 ± 0.0 0.43 ± 0.20 a 0.0 ± 0.0<br />

PQ 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0 3.0 ± 0.0 a<br />

PQ + DEX 0.67 ± 0.17 c 0.0 ± 0.0 0.44 ± 0.18 c 3.0 ± 0.0 b<br />

Kidney Control 0.22 ± 0.15 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0<br />

DEX 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0 0.0 ± 0.0<br />

PQ 1.00 ± 0.19 a 0.57 ± 0.20 a 0.88 ± 0.23 a 0.50 ± 0.19 a<br />

PQ + DEX 0.91 ± 0.21 b 0.40 ± 0.16 b 1.22 ± 0.22 b 0.56 ± 0.18 b<br />

Values are given as mean ± S.E.M. (n = 3).<br />

a p < 0.05 vs. control group.<br />

b p < 0.05 vs. DEX group.<br />

c p < 0.05 vs. PQ group.


78 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

pseudopodia emissions and degranulation with platelets<br />

aggregation, according to Ahnadi et al., 2003) and polymorphonuclear<br />

cells inside <strong>the</strong> capillaries were noticed.<br />

The majority <strong>of</strong> pneumocytes showed, at least, one ultrastructural<br />

abnormality, mitochondrial swelling being <strong>the</strong><br />

most frequent alteration. In <strong>the</strong> PQ + DEX group, comparatively<br />

to PQ group, <strong>the</strong> occurrence <strong>of</strong> <strong>the</strong> above<br />

referred alterations were drastically attenuated, particularly<br />

<strong>the</strong> amount <strong>of</strong> phagocytes observed in interstitial<br />

space or within capillaries neighboring endo<strong>the</strong>lial cells.<br />

Despite <strong>the</strong> existence <strong>of</strong> several pneumocytes with mitochondrial<br />

swelling and evidences <strong>of</strong> interstitial edema,<br />

<strong>the</strong> exuberance <strong>of</strong> those signals and <strong>the</strong> ratio <strong>of</strong> affected<br />

cells were drastically attenuated in PQ + DEX animals.<br />

Fur<strong>the</strong>rmore, comparing to <strong>the</strong> PQ group, <strong>the</strong> vascular<br />

congestion and <strong>the</strong> alveolar collapse were not so noticeable<br />

in PQ + DEX animals.<br />

Liver—major qualitative structural and ultrastructural<br />

alterations are depicted in Fig. 2. The respective<br />

semi-quantitative analysis is shown in Table 1. AtLM,<br />

animals from control and DEX groups exhibited a preserved<br />

histological structure. However, a slight cytoplasmic<br />

vacuolization identified at TEM as lipid droplets<br />

affecting few hepatocytes was observed in both groups.<br />

Animals from PQ group evidenced drastic morphological<br />

alterations, <strong>the</strong> closest hepatocytes to arterioles being<br />

<strong>the</strong> most affected by <strong>the</strong> cellular degeneration parameters.<br />

A wide cytoplasmic vacuolization resulting from<br />

intracellular edema and lipid accumulation was observed<br />

in <strong>the</strong>se animals. Moreover, extent confluent coagulative<br />

necrotic areas and several leukocytes inside sinusoids<br />

were also present. The inclusion <strong>of</strong> DEX in <strong>the</strong> experimental<br />

procedure (PQ + DEX group) attenuated <strong>the</strong><br />

severity and <strong>the</strong> incidence <strong>of</strong> <strong>the</strong> above referred injuries<br />

despite <strong>the</strong> marked hepatocyte vacuolization adjacent to<br />

portal triads.<br />

Spleen—major qualitative structural and ultrastructural<br />

alterations are depicted in Fig. 3. The respective<br />

semi-quantitative analysis is shown in Table 1. Succinctly,<br />

at LM, animals from control group showed normal<br />

splenic architecture consisting <strong>of</strong> areas <strong>of</strong> white<br />

and red pulp in equilibrated proportions. At <strong>the</strong> periphery<br />

<strong>of</strong> <strong>the</strong> periarteriolar lymphocyte sheath (PALS) we<br />

observed multinuclear macrophages. The treatment with<br />

DEX (DEX group) resulted in <strong>the</strong> disappearance <strong>of</strong> <strong>the</strong><br />

white pulp and PALS with a consequent reduction in cellular<br />

density. PQ group evidenced an apparent normal<br />

histological structure, but with lymphocytes showing<br />

clear signs <strong>of</strong> toxicity namely edema <strong>of</strong> <strong>the</strong> endoplasmic<br />

reticulum and mitochondrial swelling <strong>of</strong> <strong>the</strong> reticular<br />

and endo<strong>the</strong>lial cells. Activated platelets were also seen<br />

within sinusoids. Histological alterations observed in<br />

Fig. 2. Optical (above) and electron (below) micrographs from liver <strong>of</strong> control (A and E), dexamethasone (B and F), <strong>paraquat</strong> (C and G) and<br />

<strong>paraquat</strong> plus dexamethasone (D and H) groups. A, B, E and F evidenced a normal structure and ultrastructure with some cytoplasmic lipid droplets<br />

(blue arrows); C and G depict an extent necrotic area (*), with <strong>the</strong> presence <strong>of</strong> <strong>organ</strong>elles (red arrows) in interstitial space and lipid droplets in<br />

cytoplasm; D and H shown a diffuse and confluent cytoplasmic vacuolization, suggestive <strong>of</strong> intracellular edema (green arrows). (For interpretation<br />

<strong>of</strong> <strong>the</strong> references to color in this figure legend, <strong>the</strong> reader is referred to <strong>the</strong> web version <strong>of</strong> this article.)


R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 79<br />

Fig. 3. Optical (above) and electron (below) micrographs from spleen <strong>of</strong> control (A and E), dexamethasone (B and F), <strong>paraquat</strong> (C and G) and<br />

<strong>paraquat</strong> plus dexamethasone (D and H) groups. White pulp (#), red pulp (*) and multinuclear macrophages (blue arrows) are observed in A; two<br />

nuclei <strong>of</strong> macrophage are shown in E (blue arrow); B and F evidenced, respectively, <strong>the</strong> disappearance <strong>of</strong> <strong>the</strong> white pulp and <strong>the</strong> ultrastructure <strong>of</strong><br />

reticular connective tissue; C depict <strong>the</strong> a normal white and red pulp and G shows a macrophage, a plasmocyte cell (red arrow) and a lymphocyte<br />

with enlargement <strong>of</strong> nuclear cisterns (green arrow); D and H show <strong>the</strong> disappearance <strong>of</strong> white pulp and a damaged reticular cell (pink arrow). (For<br />

interpretation <strong>of</strong> <strong>the</strong> references to color in this figure legend, <strong>the</strong> reader is referred to <strong>the</strong> web version <strong>of</strong> this article.)<br />

<strong>the</strong> PQ + DEX group were similar to those observed in<br />

<strong>the</strong> two previous groups, although with less activated<br />

platelets and less endo<strong>the</strong>lial cells exhibiting lesions in<br />

comparison to only PQ-exposed group.<br />

Kidney—major qualitative structural and ultrastructural<br />

alterations are depicted in Fig. 4. The respective<br />

semi-quantitative analysis is shown in Table 1. A regular<br />

proximal and distal tubular structure as well as a<br />

normal glomerular architecture was registered in <strong>the</strong> control<br />

and DEX groups, although several proximal tubular<br />

cells with mitochondrial swelling were observed in<br />

DEX group. PQ group evidenced marked tubular lesions,<br />

particularly notorious in <strong>the</strong> proximal tubule with confluent<br />

areas <strong>of</strong> vacuolated cells, apparently resulting<br />

from mitochondrial swelling, and coagulative necrosis<br />

with a tubular cell loss. Distal tubule was slightly<br />

affected. The glomeruli showed moderate alterations<br />

affecting endo<strong>the</strong>lial cells. Several infiltrative cells were<br />

also observed in interstitial space. In opposition to lung<br />

and liver <strong>the</strong> above referred alterations were not ameliorated<br />

by DEX.<br />

3.2. Relative <strong>organ</strong> weight<br />

Data on <strong>the</strong> relative weights <strong>of</strong> lung (RLW), liver<br />

(RLiW), spleen (RSW) and kidney (RKW) are present in<br />

Table 2. In comparison to <strong>the</strong> control group, animals from<br />

PQ group showed a significant RLW increase (p < 0.05),<br />

whereas RLW <strong>of</strong> DEX-post-treated animals (PQ + DEX<br />

group) were near to <strong>the</strong> control. RLiW in rats from DEX,<br />

PQ and PQ + DEX groups were comparable to controls.<br />

Animals from <strong>the</strong> DEX group showed a significant RSW<br />

decrease (p < 0.01) relatively to control group. RSW in<br />

PQ-exposed animals was similar to that <strong>of</strong> control group.<br />

The inclusion <strong>of</strong> DEX in <strong>the</strong> PQ treatment (PQ + DEX<br />

group) caused a decrease <strong>of</strong> RSW (p < 0.01 versus control<br />

group), comparable to <strong>the</strong> decrease observed in <strong>the</strong><br />

DEX group. Rats exposed to PQ exhibited an increase<br />

<strong>of</strong> <strong>the</strong> RKW (p < 0.05) that was not reverted by DEX<br />

administration (p < 0.05 versus control group).<br />

3.3. LPO and carbonyl groups content<br />

As shown in Table 2, animals from PQ group exhibited<br />

a significant increase <strong>of</strong> <strong>the</strong> LPO in lungs comparing<br />

to control group (p < 0.001). On <strong>the</strong> o<strong>the</strong>r hand, DEX<br />

administration reverted this parameter down to near control<br />

levels. Analogous results were obtained for carbonyl<br />

groups content (Table 2). Similar pr<strong>of</strong>iles were observed<br />

for carbonyl groups in <strong>the</strong> hepatic tissue. LPO and carbonyl<br />

groups increased in <strong>the</strong> kidney <strong>of</strong> PQ-exposed rats<br />

relatively to control group (p < 0.05, respectively). In <strong>the</strong>


80 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

Fig. 4. Optical (above) and electron (below) micrographs from kidney <strong>of</strong> control (A and E), dexamethasone (B and F), <strong>paraquat</strong> (C and G) and<br />

<strong>paraquat</strong> plus dexamethasone (D and H) groups. A and B evidenced a glomerular and tubular normal structure while E and F depict, respectively,<br />

a podocyte attached to basement membrane (#) and several proximal tubular cells with mitochondrial swelling (blue arrows); in C and D it could<br />

be observed a necrotic area (*) with <strong>the</strong> presence <strong>of</strong> cell <strong>organ</strong>elles in interstitial space (red arrows); G and H show a cytoplasmic vacuolization <strong>of</strong><br />

proximal tubular cells resulting from mitochondrial swelling and from <strong>the</strong> presence <strong>of</strong> large cytoplasmic vacuoles (green arrows). (For interpretation<br />

<strong>of</strong> <strong>the</strong> references to color in this figure legend, <strong>the</strong> reader is referred to <strong>the</strong> web version <strong>of</strong> this article.)<br />

spleen, LPO increased in <strong>the</strong> PQ group in comparison to<br />

<strong>the</strong> control group (p < 0.05). Noteworthy are <strong>the</strong> increase<br />

<strong>of</strong> LPO and carbonyl groups observed in <strong>the</strong> spleen <strong>of</strong><br />

DEX group relatively to control group (p < 0.01 and 0.05,<br />

respectively), as well as <strong>the</strong> lack <strong>of</strong> protective effect <strong>of</strong><br />

DEX in liver LPO and kidney LPO and carbonyl groups.<br />

3.4. MPO activity<br />

Aliquots <strong>of</strong> rat <strong>organ</strong> samples were assayed for <strong>the</strong><br />

activity <strong>of</strong> MPO, which is an index <strong>of</strong> neutrophils sequestration,<br />

26 h after exposure to PQ. As shown in Table 2,<br />

lung MPO activity <strong>of</strong> <strong>the</strong> PQ-exposed animals was significantly<br />

higher (with a p < 0.05) than in rats from control<br />

group. The post-treatment with DEX, completely<br />

prevented <strong>the</strong> increase <strong>of</strong> MPO activity. Liver MPO<br />

activity revealed similar results to those observed in<br />

<strong>the</strong> lung. In <strong>the</strong> kidney, an increase <strong>of</strong> MPO activity in<br />

<strong>the</strong> PQ group in relation to control group was observed<br />

(p < 0.05), but DEX did not provide any protective effect.<br />

MPO expression was also not modified in spleen after<br />

PQ exposure, although an increase <strong>of</strong> its activity was<br />

observed in <strong>the</strong> PQ + DEX relatively to control and PQ<br />

group (p < 0.01, respectively). The results also showed<br />

that DEX led to an increase <strong>of</strong> MPO activity in <strong>the</strong> spleen,<br />

comparatively to <strong>the</strong> control group (p < 0.01).<br />

3.5. Quantification <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> rat lung,<br />

kidney, spleen and liver<br />

The PQ lung concentration <strong>of</strong> <strong>the</strong> PQ group was<br />

0.127 ± 0.010 [(mean ± S.E.M.), �g/mg protein]. Animals<br />

post-treated with DEX evidenced a significant<br />

decrease in PQ lung concentration, down to<br />

0.062 ± 0.008 (p < 0.05) (Table 3). The PQ concentration<br />

in kidney, spleen and liver <strong>of</strong> <strong>the</strong> PQ group did<br />

not evidence any significant difference comparatively to<br />

PQ + DEX group (Table 3).<br />

3.6. Urinary NAG<br />

NAG urinary excretion was significantly increased<br />

26 h (Fig. 5) after exposure <strong>of</strong> rats to PQ comparatively<br />

to <strong>the</strong> control group. The administration <strong>of</strong> DEX<br />

(PQ + DEX group) resulted in a fur<strong>the</strong>r increase <strong>of</strong> NAG<br />

urinary excretion.<br />

3.7. Effect <strong>of</strong> dexamethasone and verapamil on <strong>the</strong><br />

survival <strong>of</strong> <strong>paraquat</strong>-exposed rats and o<strong>the</strong>r<br />

observations<br />

Diarrhoea, piloerection, weight loss, anorexia, adipsia,<br />

hyperpnea, dyspnea, tachycardia and a red drainage


R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 81<br />

Table 2<br />

Relative <strong>organ</strong>s weight (ROW) and toxicological parameters <strong>of</strong> <strong>the</strong> control, dexamethasone (DEX), <strong>paraquat</strong> (PQ) and <strong>paraquat</strong> + dexamethasone<br />

(PQ + DEX)<br />

Organ Group Evaluated parameter<br />

ROW TBARS (nmol MDA/mg<br />

protein)<br />

Carbonyl groups<br />

(nmol/mg protein)<br />

MPO (U/g protein)<br />

Lung Control 0.37 ± 0.01 0.192 ± 0.015 1.860 ± 0.098 21.569 ± 2.232<br />

DEX 0.36 ± 0.02 0.200 ± 0.020 1.789 ± 0.064 20.067 ± 1.989<br />

PQ 0.43 ± 0.02 a,b 0.485 ± 0.033 aaa,bbb 2.254 ± 0.135 a,b 29.143 ± 1.915 a,b<br />

PQ + DEX 0.36 ± 0.02 0.274 ± 0.039 cc 2.013 ± 0.193 20.621 ± 2.565 c<br />

Liver Control 4.14 ± 0.11 0.135 ± 0.037 2.020 ± 0.301 11.854 ± 0.549<br />

DEX 4.01 ± 0.10 0.130 ± 0.040 1.969 ± 0.298 11.278 ± 0.860<br />

PQ 3.90 ± 0.21 0.200 ± 0.055 a,b 2.921 ± 0.183 a,b 15.875 ± 0.975 a,b<br />

PQ + DEX 4.08 ± 0.15 0.182 ± 0.082 2.342 ± 0.202 12.984 ± 1.034 c<br />

Spleen Control 0.25 ± 0.01 0.271 ± 0.026 0.582 ± 0.041 75.777 ± 1.298<br />

DEX 0.18 ± 0.02 aa 0.441 ± 0.031 aa 0.652 ± 0.053 a 80.201 ± 2.890 a<br />

PQ 0.27 ± 0.03 0.322 ± 0.039 a,b 0.594 ± 0.072 b 73.532 ± 1.927 b<br />

PQ + DEX 0.18 ± 0.01 aa,cc 0.537 ± 0.105 aa,cc 0.663 ± 0.091 a,c 82.939 ± 1.282 aa,cc<br />

Kidney Control 0.51 ± 0.01 0.712 ± 0.064 3.110 ± 0.191 6.447 ± 0.204<br />

DEX 0.52 ± 0.01 0.734 ± 0.039 3.087 ± 0.143 6.767 ± 0.239<br />

PQ 0.56 ± 0.02 a,b 0.925 ± 0.169 a,b 3.932 ± 0.129 a,b 8.855 ± 0.586 a<br />

PQ + DEX 0.57 ± 0.01 a,b 1.056 ± 0.223 aa,bb 3.987 ± 0.152 a,b 8.448 ± 0.347 a<br />

Values are given as mean ± S.E.M. (n = 5).<br />

a p < 0.05 vs. control group.<br />

aa p < 0.01 vs. control group.<br />

aaa p < 0.001 vs. control group.<br />

b p < 0.05 vs. DEX group.<br />

bb p < 0.01 vs. DEX group.<br />

bbb p < 0.001 vs. DEX group.<br />

c p < 0.05 vs. PQ group.<br />

cc p < 0.01 vs. PQ group.<br />

around <strong>the</strong> mouth, eyes and nose were present especially<br />

in animals exposed to PQ and PQ + VER + DEX<br />

during <strong>the</strong> first 48 h. During <strong>the</strong> same experimental<br />

period, rats belonging to PQ + VER + DEX group did<br />

not ingest any amount <strong>of</strong> water and only a few milliliters<br />

were ingested by rats <strong>of</strong> PQ group. Deep breathing<br />

was observed and <strong>the</strong> thorax was sunken in <strong>the</strong> animals<br />

from PQ and PQ + VER + DEX groups in contrast<br />

to those belonging to control, DEX or PQ + DEX-<br />

Table 3<br />

PQ lung, kidney, spleen and liver concentration in <strong>the</strong> <strong>paraquat</strong> (PQ)<br />

and <strong>paraquat</strong> plus dexamethasone (PQ + DEX) groups<br />

Organ PQ levels (�g/mg protein)<br />

PQ PQ + DEX<br />

Lung 0.129 ± 0.062 0.062 ± 0.008 a<br />

Kidney 0.029 ± 0.005 0.033 ± 0.011<br />

Spleen 0.015 ± 0.005 0.016 ± 0.003<br />

Liver 0.008 ± 0.001 0.008 ± 0.001<br />

Values are given as mean ± S.E.M. (n = 5).<br />

a p < 0.05 vs. PQ group.<br />

treated groups. Rats exposed only to PQ (PQ group)<br />

displayed approximately, 25 and 100% <strong>of</strong> mortality<br />

by <strong>the</strong> 2nd and 6th day, respectively (Fig. 6). Hundred<br />

percent <strong>of</strong> mortality was observed by <strong>the</strong> 4th day<br />

in <strong>the</strong> group PQ + VER + DEX. Post-treatment <strong>of</strong> PQ-<br />

Fig. 5. Urinary N-acetyl-�-d-glucosaminidase (NAG) activity in <strong>the</strong><br />

control, <strong>paraquat</strong> (PQ) and <strong>paraquat</strong> plus dexamethasone (PQ + DEX)<br />

groups. Values are given as mean ± S.E.M. (n = 8). aa p < 0.01 and<br />

aaa p < 0.001 vs. control.


82 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

Fig. 6. Percentage <strong>of</strong> rat survival in <strong>the</strong> control, <strong>paraquat</strong> (PQ),<br />

<strong>paraquat</strong> plus dexamethasone (PQ + DEX) groups and <strong>paraquat</strong> plus<br />

verapamil and dexamethasone (PQ + VER + DEX) groups. c p < 0.05<br />

vs. PQ group.<br />

exposed rats with DEX (PQ + DEX group) resulted in<br />

a significant enhancement <strong>of</strong> <strong>the</strong> survival time (50%<br />

<strong>of</strong> survival at 10th day, p < 0.05). Logrank test showed<br />

significant differences between <strong>the</strong> survival curves <strong>of</strong><br />

PQ versus PQ + DEX and PQ + VER + DEX (p < 0.05,<br />

respectively).<br />

4. Discussion<br />

The main objective <strong>of</strong> <strong>the</strong> present study was to assess<br />

<strong>the</strong> effect <strong>of</strong> DEX in PQ-exposed rats, in <strong>the</strong> lung as well<br />

as in o<strong>the</strong>r <strong>organ</strong>s and systems, but also to support our<br />

hypo<strong>the</strong>sis that <strong>the</strong> protective effect <strong>of</strong> DEX against PQ<br />

toxicity might, at least, be partially mediated by P-gp<br />

functionality. Lung was undoubtedly <strong>the</strong> most affected<br />

<strong>organ</strong>, which is in accordance to <strong>the</strong> accumulation <strong>of</strong><br />

PQ in this <strong>organ</strong> through a highly developed polyamine<br />

uptake system (Dinis-Oliveira et al., 2006c; Nemery et<br />

al., 1987; Rannels et al., 1989). As we showed previously<br />

(Dinis-Oliveira et al., 2006b) <strong>the</strong> high dose <strong>of</strong> DEX used<br />

in <strong>the</strong> present study leads to a stunning increase <strong>of</strong> lung<br />

P-gp expression. The involvement <strong>of</strong> P-gp in decreasing<br />

PQ-lung concentration is evidenced by <strong>the</strong> lowering<br />

effect on PQ levels mediated by DEX and <strong>the</strong> respective<br />

inhibition by VER, a competitive inhibitor <strong>of</strong> this transporter.<br />

In this study, we observed that this <strong>the</strong>rapeutic<br />

approach conducts to an increase in <strong>the</strong> survival rate <strong>of</strong><br />

animals belonging to PQ + DEX group in comparison to<br />

animals only exposed to PQ.<br />

The increase <strong>of</strong> <strong>the</strong> RLW (Table 2) and <strong>the</strong> presence <strong>of</strong><br />

interstitial edema evidenced by histopathological analysis<br />

(Fig. 1) confirmed that PQ <strong>induced</strong> lung edema,<br />

an effect that was drastically attenuated in PQ + DEX<br />

treated animals. The cellular damage mediated by PQ<br />

is essentially due to its redox-cycle leading to continuous<br />

superoxide radicals (O2 •− ) production (Bus et al.,<br />

1974). This <strong>the</strong>n sets in <strong>the</strong> well-known cascade leading<br />

to generation <strong>of</strong> <strong>the</strong> hydroxyl radical (HO • )(Youngman<br />

and Elstner, 1981), which has been implicated in <strong>the</strong> initiation<br />

<strong>of</strong> membrane injury by lipid peroxidation (LPO)<br />

during <strong>the</strong> exposure to PQ in vitro (Bus et al., 1975)<br />

as well as in vivo (Chen and Lua, 2000). Besides,<br />

<strong>research</strong>ers have been suggesting <strong>the</strong> hypo<strong>the</strong>sis <strong>of</strong> cytotoxicity<br />

via mitochondrial dysfunction caused by PQ<br />

(Dinis-Oliveira et al., 2006d; Fukushima et al., 1994).<br />

We have previously demonstrated that <strong>the</strong> increase <strong>of</strong><br />

LPO <strong>induced</strong> by PQ-exposure was significantly reduced<br />

by DEX (Dinis-Oliveira et al., 2006b). The present study<br />

corroborates those results. Besides lipids, ROS are also<br />

known to oxidatively modify DNA, carbohydrates and<br />

proteins. Fragmentation <strong>of</strong> polypeptide chains, increased<br />

sensitivity to denaturation, formation <strong>of</strong> protein–protein<br />

cross-linkages as well as modification <strong>of</strong> amino acid side<br />

chains to hydroxyl or carbonyl derivatives are possible<br />

outcomes <strong>of</strong> protein oxidative reactions (Dean et al.,<br />

1997). Accordingly, it was also shown that PQ administration<br />

increased <strong>the</strong> carbonyl groups content in lung<br />

and in accordance to previous results (Dinis-Oliveira et<br />

al., 2006b), DEX protected against PQ-<strong>induced</strong> increase<br />

<strong>of</strong> carbonyl groups content. In <strong>the</strong> present study, <strong>the</strong><br />

histopathological findings confirmed that PQ <strong>induced</strong><br />

marked alterations to <strong>the</strong> normal pattern <strong>of</strong> lung, with<br />

majority <strong>of</strong> pneumocytes showing, at least, one ultrastructural<br />

abnormality, mitochondrial swelling being <strong>the</strong><br />

most frequent alteration. These morphological evidences<br />

<strong>of</strong> cellular aggression were again attenuated by DEXtreatment,<br />

results evidenced by qualitative and quantitative<br />

analysis <strong>of</strong> <strong>the</strong> morphological injury (Table 1<br />

and Fig. 1). In addition, <strong>the</strong> reduced amount <strong>of</strong> activated<br />

platelets within <strong>the</strong> capillaries observed in DEXtreated<br />

animals might be interpreted as a consequence <strong>of</strong><br />

endo<strong>the</strong>lial cells protection against PQ-toxicity.<br />

Considering <strong>the</strong> liver, RLiW measurements did not<br />

reveal any difference between <strong>the</strong> experimental groups,<br />

although a wide cytoplasmic vacuolization was observed<br />

in <strong>the</strong> PQ group. Besides <strong>the</strong> low PQ concentrations<br />

quantified in this <strong>organ</strong>, necrotic zones and tissue dis<strong>organ</strong>ization<br />

were notorious in PQ-exposed animals,<br />

particularly surrounding centrilobular region. The more<br />

susceptibility <strong>of</strong> this region can be explained by its higher<br />

concentration in NADPH-cytochrome P-450 reductase<br />

(Jungermann and Kietzmann, 1996), essential to PQ<br />

redox-cycle (Clejan and Cederbaum, 1989) and consequent<br />

oxidative stress propagation. Moreover, although<br />

<strong>the</strong> PQ elimination occurs mainly through kidneys, <strong>the</strong><br />

biliar excretion (Dinis-Oliveira et al., 2006b; Hughes<br />

et al., 1973) may also have contributed to <strong>the</strong> discrepancy<br />

observed between PQ concentrations and <strong>the</strong> extent<br />

<strong>of</strong> <strong>the</strong> lesions in this <strong>organ</strong>. Similar histopathological


esults were also previously described in animals and<br />

humans (Burk et al., 1980; Parkinson, 1980). However,<br />

our results showed, for <strong>the</strong> first time, that DEX significantly<br />

reduced signs <strong>of</strong> cell degeneration, interstitial<br />

inflammatory cell infiltration, necrotic zones and tissue<br />

dis<strong>organ</strong>ization in <strong>the</strong> liver <strong>of</strong> PQ-exposed rats. Additionally,<br />

and as it was previously observed in lung, hepatic<br />

alterations in <strong>the</strong> LPO and protein carbonylation,<br />

were correlated with <strong>the</strong> extent <strong>of</strong> histological damage.<br />

Regarding <strong>the</strong> spleen, no significant changes were<br />

observed for RSW between control and PQ group. Taking<br />

<strong>into</strong> account that one <strong>of</strong> <strong>the</strong> major functions <strong>of</strong> <strong>the</strong><br />

spleen is to remove damaged erythrocytes, and since PQ<br />

proved to damage erythrocytes by altering its antioxidant<br />

status (Hernandez et al., 2005), it is expected that<br />

injured erythrocytes will be ultimately scavenged by <strong>the</strong><br />

spleen, generating ROS and subsequent tissue injury.<br />

In <strong>the</strong> present study, LPO increased in <strong>the</strong> spleen <strong>of</strong><br />

PQ-exposed rats in relation to control group. According<br />

to that, qualitative and quantitative analysis <strong>of</strong> <strong>the</strong><br />

morphological injury revealed tissue dis<strong>organ</strong>ization <strong>of</strong><br />

PQ-exposed rats, mitochondrial swelling <strong>of</strong> <strong>the</strong> reticular<br />

and endo<strong>the</strong>lial cells being <strong>the</strong> most significant alteration<br />

observed in this group. Interestingly, a decrease <strong>of</strong> <strong>the</strong><br />

RSW was observed in <strong>the</strong> PQ + DEX group, which may<br />

be due to a consequence <strong>of</strong> spleen atrophy caused by<br />

DEX (Orzechowski et al., 2002). Our results also showed<br />

that DEX, by itself, reduced RSW and caused <strong>the</strong> disappearance<br />

<strong>of</strong> <strong>the</strong> white pulp (DEX group). Since white<br />

pulp reflects T-cell mass, such effect probably corresponds<br />

to <strong>the</strong> immunosuppressive effectiveness <strong>of</strong> DEX,<br />

that it is in accordance with current <strong>the</strong>rapeutic guidelines<br />

to prevent pulmonary fibrosis (Mason et al., 1999).<br />

The lysosomal enzyme NAG is generally regarded<br />

as an indicator <strong>of</strong> renal tubular dysfunction and disease<br />

(Price, 1982). In this work, <strong>the</strong> increase <strong>of</strong> NAG urinary<br />

excretion <strong>induced</strong> by PQ was accompanied by an<br />

increase in LPO, protein carbonylation, proximal tubular<br />

damage, coagulative necrosis and tubular cell loss.<br />

Similar results were also documented by Murray and<br />

Gibson (1972). Urine is <strong>the</strong> main excretion via <strong>of</strong> PQ<br />

and this toxicity seems to result from intracellular redoxcycle<br />

generated by PQ in proximal tubules (Lock and<br />

Ishmael, 1979). Noteworthy were <strong>the</strong> results observed in<br />

<strong>the</strong> PQ + DEX exposed rats. Unexpectedly, DEX aggravated<br />

PQ-<strong>induced</strong> kidney toxicity, leading to an increase<br />

<strong>of</strong> LPO and protein carbonylation. In accordance, NAG<br />

urinary excretion steeply increased and RKW did not<br />

ameliorate in <strong>the</strong> PQ + DEX group relatively to PQ<br />

group. This lack <strong>of</strong> kidney protection caused by DEX in<br />

PQ-exposed rats might be <strong>the</strong> consequence <strong>of</strong> <strong>the</strong> <strong>organ</strong><br />

specificity regarding <strong>the</strong> P-gp expression as consequence<br />

R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 83<br />

<strong>of</strong> DEX treatment. Indeed, while DEX increases P-gp<br />

expression in liver and lung, it has an opposite effect in<br />

<strong>the</strong> kidney (Demeule et al., 1999). In this way, less P-gp<br />

expression by DEX in <strong>the</strong> kidneys will cause an extended<br />

presence <strong>of</strong> PQ in <strong>the</strong> proximal tubules and consequently,<br />

more damage and urinary NAG release will take place.<br />

Never<strong>the</strong>less, although a tendency for higher PQ levels<br />

(∼11%) was observed in <strong>the</strong> kidney <strong>of</strong> PQ + DEX group,<br />

comparatively to <strong>the</strong> PQ group, <strong>the</strong> non-significance <strong>of</strong><br />

this result indicates that o<strong>the</strong>r <strong>mechanisms</strong> are probably<br />

involved.<br />

It should be considered that <strong>the</strong> observed DEX protective<br />

effects against PQ-<strong>induced</strong> lung and liver toxicity<br />

may also result from its anti-inflammatory effects.<br />

Indeed, according to Hybertson et al. (1995), <strong>the</strong> toxicity<br />

provoked by PQ is assumed to be associated<br />

with <strong>the</strong> activation <strong>of</strong> neutrophils. Fur<strong>the</strong>rmore, various<br />

inflammatory mediators have been found to be<br />

increased in <strong>the</strong> alveolar space during <strong>the</strong> early phase<br />

<strong>of</strong> ARDS, including tumor necrosis factor-alpha (TNF-<br />

�), interleukin-1�, interleukin-6 and chemokines (Pugin<br />

et al., 1999), which stimulate <strong>the</strong> infiltration <strong>of</strong> polymorphonuclear<br />

leukocytes (PMN) <strong>into</strong> <strong>the</strong> lungs. DEX<br />

has been shown to decrease TNF-� concentrations in<br />

<strong>the</strong> bronchoalveolar lavage fluid <strong>of</strong> PQ treated rats<br />

(Chen et al., 2001). DEX presents also an inhibitory<br />

effect on ROS production by macrophages and neutrophils<br />

(Maridonneau-Parini et al., 1989). Since MPO is<br />

located within <strong>the</strong> primary azurophil granules <strong>of</strong> PMN,<br />

its activity indirectly reflects PMN infiltration through<br />

<strong>the</strong> <strong>organ</strong>s (Schultz and Kaminker, 1962) during <strong>the</strong><br />

inflammatory reaction. As expected, our results showed<br />

that MPO activity is markedly elevated in lung, liver and<br />

kidney <strong>of</strong> PQ-exposed animals. Our histopathological<br />

results confirmed <strong>the</strong> widespread neutrophils infiltration<br />

in <strong>the</strong>se <strong>organ</strong>s. DEX administration caused a significant<br />

decrease <strong>of</strong> <strong>the</strong> interstitial inflammatory cell infiltration<br />

score, in lung and liver, <strong>of</strong> animals exposed to PQ.<br />

Despite <strong>the</strong> beneficial effects observed in <strong>the</strong> lungs<br />

and liver <strong>of</strong> PQ-<strong>into</strong>xicated rats treated with DEX, this<br />

protection appears not be achieved in <strong>the</strong> kidney and<br />

spleen. In fact, this study confirmed once more that PQ<br />

poisoning is an extreme frustrating condition to manage.<br />

In attempt to verify <strong>the</strong> contribution <strong>of</strong> <strong>the</strong>se apparent<br />

contradictory results to <strong>the</strong> final outcome we assessed<br />

<strong>the</strong> survival rate <strong>of</strong> this approach. If still some doubts<br />

existed, DEX proved to increase <strong>the</strong> survival rate by shifting<br />

<strong>the</strong> time course <strong>of</strong> deaths (Fig. 6). Giving credit to this<br />

protection, VER showed <strong>the</strong> tremendous contribution <strong>of</strong><br />

P-gp functionality to <strong>the</strong> final outcome. Indeed rats that<br />

received VER prior to DEX (PQ + VER + DEX group)<br />

died within 48 h, faster than rats that were only PQ-


84 R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85<br />

exposed (PQ group). On <strong>the</strong> o<strong>the</strong>r hand, only PQ + DEX<br />

group had animals that survived beyond <strong>the</strong> 5th day, with<br />

50% rats remaining alive by <strong>the</strong> 10th day. It is important<br />

to focus that this improvement in <strong>the</strong> survival rate was<br />

obtained with only a single dose <strong>of</strong> DEX. It might be<br />

supposed that repetitive DEX <strong>the</strong>rapy could extend survival<br />

time and allow a lung transplant to be performed<br />

in <strong>the</strong> PQ-poisoned patients. Following <strong>the</strong>se encouraging<br />

results, fur<strong>the</strong>r studies are needed to clarify <strong>the</strong>se<br />

protective effects.<br />

Acknowledgement<br />

Ricardo Dinis acknowledges FCT for his Ph.D. grant<br />

(SFRH/BD/13707/2003).<br />

References<br />

Ahnadi, C.E., Chapman, E.S., Lepine, M., Okrongly, D., Pujol-Moix,<br />

N., Hernandez, A., Boughrassa, F., Grant, A.M., 2003. Assessment<br />

<strong>of</strong> platelet activation in several different anticoagulants by<br />

<strong>the</strong> Advia 120 Hematology System, fluorescence flow cytometry,<br />

and electron microscopy. Thromb. Haemost. 90, 940–948.<br />

Akahori, F., Masaoka, T., Matsushiro, S., Arishima, K., Arai, S.,<br />

Yamamoto, M., Eguchi, Y., 1987. Quantifiable morphologic evaluation<br />

<strong>of</strong> <strong>paraquat</strong> pulmonary toxicity in rats. Vet. Hum. Toxicol.<br />

29, 1–7.<br />

Andrews, P.C., Krinsky, N.I., 1982. Quantitative determination <strong>of</strong><br />

myeloperoxidase using tetramethylbenzidine as substrate. Anal.<br />

Biochem. 127, 346–350.<br />

Ascensao, A., Magalhaes, J., Soares, J.M., Ferreira, R., Neuparth, M.J.,<br />

Marques, F., Oliveira, P.J., Duarte, J.A., 2005. Moderate endurance<br />

training prevents doxorubicin-<strong>induced</strong> in vivo mitochondriopathy<br />

and reduces <strong>the</strong> development <strong>of</strong> cardiac apoptosis. Am. J. Physiol.<br />

Heart Circ. Physiol. 289, H722–H731.<br />

Bismuth, C., Garnier, R., Baud, F.J., Muszynski, J., Keyes, C., 1990.<br />

Paraquat poisoning. An overview <strong>of</strong> <strong>the</strong> current status. Drug Saf.,<br />

243–251.<br />

Buege, J.A., Aust, S.D., 1978. Microsomal lipid peroxidation. Methods<br />

Enzymol. 52, 302–310.<br />

Burk, R.F., Lawrence, R.A., Lane, J.M., 1980. Liver necrosis and lipid<br />

peroxidation in <strong>the</strong> rat as result <strong>of</strong> <strong>paraquat</strong> and diquat administration.<br />

Effect <strong>of</strong> selenium deficiency. J. Clin. Invest. 65, 1024–1031.<br />

Bus, J.S., Aust, S.D., Gibson, J.E., 1974. Superoxide- and singlet<br />

oxygen-catalyzed lipid peroxidation as a possible mechanism for<br />

<strong>paraquat</strong> (methyl viologen) toxicity. Biochem. Biophys. Res. Commun.<br />

58, 749–755.<br />

Bus, J.S., Aust, S.D., Gibson, J.E., 1975. Lipid peroxidation: a possible<br />

mechanism for <strong>paraquat</strong> toxicity. Res. Commun. Chem. Pathol.<br />

Pharmacol. 11, 31–38.<br />

Carvalho, F., Fernandes, E., Remiao, F., Bastos, M.L., 1999. Effect <strong>of</strong> damphetamine<br />

repeated administration on rat antioxidant defences.<br />

Arch. Toxicol. 73, 83–89.<br />

Chatterjee, P.K., Cuzzocrea, S., Brown, P.A., Zacharowski, K., Stewart,<br />

K.N., Mota-Filipe, H., Thiemermann, C., 2000. Tempol, a<br />

membrane-permeable radical scavenger, reduces oxidant stressmediated<br />

renal dysfunction and injury in <strong>the</strong> rat. Kidney Int. 58,<br />

658–673.<br />

Chen, C.M., Lua, A.C., 2000. Lung toxicity <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> rat. J.<br />

Toxicol. Environ. Health A 60, 477–487.<br />

Chen, C.M., Fang, C.L., Chang, C.H., 2001. Surfactant and corticosteroid<br />

effects on lung function in a rat model <strong>of</strong> acute lung injury.<br />

Crit. Care Med. 29, 2169–2175.<br />

Chen, G.H., Lin, J.L., Huang, Y.K., 2002. Combined methylprednisolone<br />

and dexamethasone <strong>the</strong>rapy for <strong>paraquat</strong> poisoning. Crit.<br />

Care Med. 30, 2584–2587.<br />

Chen, C.M., Wang, L.F., Su, B., Hsu, H.H., 2003. Methylprednisolone<br />

effects on oxygenation and histology in a rat model <strong>of</strong> acute lung<br />

injury. Pulm. Pharmacol. Ther. 16, 215–220.<br />

Clark, D.G., McElligott, T.F., Hurst, E.W., 1966. The toxicity <strong>of</strong><br />

<strong>paraquat</strong>. Br. J. Ind. Med. 23, 126–132.<br />

Clejan, L., Cederbaum, A.I., 1989. Synergistic interaction between<br />

NADPH-cytochrome P-450 reductase, <strong>paraquat</strong> and iron in <strong>the</strong><br />

generation <strong>of</strong> active oxygen radicals. Biochem. Pharmacol. 38,<br />

1779–1786.<br />

Dean, R.T., Fu, S., Stocker, R., Davies, M.J., 1997. Biochemistry and<br />

pathology <strong>of</strong> radical-mediated protein oxidation. Biochem. J. 324,<br />

1–18.<br />

Demeule, M., Jodoin, J., Beaulieu, E., Brossard, M., Beliveau, R., 1999.<br />

Dexamethasone modulation <strong>of</strong> multidrug transporters in normal<br />

tissues. FEBS Lett. 442, 208–214.<br />

Dinis-Oliveira, R.J., Sarmento, A., Reis, P., Amaro, A., Remião, F.,<br />

Bastos, M.L., Carvalho, F., 2006a. Acute <strong>paraquat</strong> poisoning:<br />

report <strong>of</strong> a survival case following intake <strong>of</strong> a potential lethal dose.<br />

Pediatr. Emerg. Care 22, 537–540.<br />

Dinis-Oliveira, R.J., Remião, F., Duarte, J.A., Sánchez-Navarro, A.,<br />

Bastos, M.L., Carvalho, F., 2006b. P-glycoprotein induction: an<br />

antidotal pathway for <strong>paraquat</strong>-<strong>induced</strong> lung toxicity. Free Radic.<br />

Biol. Med., doi:10.1016/j.freeradbiomed.2006.06.012.<br />

Dinis-Oliveira, R.J., Valle, M.J., Bastos, M.L., Carvalho, F., Sánchez-<br />

Navarro, A., 2006c. Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung.<br />

Influence <strong>of</strong> sodium depletion. Xenobiotica 36, 724–737.<br />

Dinis-Oliveira, R.J., Remião, F., Duarte, J.A., Sánchez-Navarro, A.,<br />

Bastos, M.L., Carvalho, F., 2006d. Paraquat exposure as an etiological<br />

factor <strong>of</strong> Parkinson’s disease. Neurotoxicology, doi:10.1016/j.<br />

neuro.2006.05.012.<br />

Duarte, J.A., Leao, A., Magalhaes, J., Ascensao, A., Bastos, M.L.,<br />

Amado, F.L., Vilarinho, L., Quelhas, D., Appell, H.J., Carvalho,<br />

F., 2005. Strenuous exercise aggravates MDMA-<strong>induced</strong> skeletal<br />

muscle damage in mice. Toxicology 206, 349–358.<br />

Fuke, C., Ameno, K., Ameno, S., Kiriu, T., Shinohara, T., Sogo, K.,<br />

Ijiri, I., 1992. A rapid, simultaneous determination <strong>of</strong> <strong>paraquat</strong> and<br />

diquat in serum and urine using second-derivative spectroscopy. J.<br />

Anal. Toxicol. 16, 214–216.<br />

Fukushima, T., Yamada, K., Hojo, N., Isobe, A., Shiwaku, K.,<br />

Yamane, Y., 1994. Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. III.<br />

The effects <strong>of</strong> acute <strong>paraquat</strong> exposure on <strong>the</strong> electron transport<br />

system in rat mitochondria. Exp. Toxicol. Pathol. 46, 437–<br />

441.<br />

Hernandez, A.F., Lopez, O., Rodrigo, L., Gil, F., Pena, G., Serrano, J.L.,<br />

Parron, T., Alvarez, J.C., Lorente, J.A., Pla, A., 2005. Changes in<br />

erythrocyte enzymes in humans long-term exposed to pesticides:<br />

influence <strong>of</strong> several markers <strong>of</strong> individual susceptibility. Toxicol.<br />

Lett. 159, 13–21.<br />

Hughes, R.D., Millburn, P., Williams, R.T., 1973. Biliary excretion <strong>of</strong><br />

some diquaternary ammonium cations in <strong>the</strong> rat, guinea pig and<br />

rabbit. Biochem. J. 136, 979–984.<br />

Hybertson, B.M., Lampey, A.S., Clarke, J.H., Koh, Y., Repine, J.E.,<br />

1995. N-acetylcysteine pretreatment attenuates <strong>paraquat</strong>-<strong>induced</strong><br />

lung leak in rats. Redox Rep. 1, 337–342.


Jungermann, K., Kietzmann, T., 1996. Zonation <strong>of</strong> parenchymal and<br />

nonparenchymal metabolism in liver. Annu. Rev. Nutr., 179–203.<br />

Levine, R.L., Williams, J.A., Stadtman, E.R., Shacter, E., 1994. Carbonyl<br />

assay for determination <strong>of</strong> oxidatively modified proteins.<br />

Methods Enzymol. 233, 346–357.<br />

Lock, E.A., Ishmael, J., 1979. The acute toxic effects <strong>of</strong> <strong>paraquat</strong> and<br />

diquat on <strong>the</strong> rat kidney. Toxicol. Appl. Pharmacol. 50, 67–76.<br />

Lowry, O.H.N., Rosebrough, N.J., Farr, A.L., Randall, R.J., 1951. Protein<br />

measurement with Folin phenol reagent. J. Biol. Chem. 193,<br />

265–275.<br />

Maridonneau-Parini, I., Errasfa, M., Russo-Marie, F., 1989. Inhibition<br />

<strong>of</strong> O2-generation by dexamethasone is mimicked by lipocortin I in<br />

alveolar macrophages. J. Clin. Invest. 83, 1936–1940.<br />

Mason, R.J., Schwarz, M.I., Hunninghake, G.W., Musson, R.A., 1999.<br />

NHLBI Workshop Summary. Pharmacological <strong>the</strong>rapy for idiopathic<br />

pulmonary fibrosis. Past, present, and future. Am. J. Respir.<br />

Crit. Care Med. 160, 1771–1777.<br />

Meduri, G.U., Belenchia, J.M., Estes, R.J., Wunderink, R.G., Torky,<br />

M.el., Leeper, K.V.J., 1991. Fibroproliferative phase <strong>of</strong> ARDS.<br />

Clinical findings and effects <strong>of</strong> corticosteroids. Chest 100,<br />

943–952.<br />

Melchiorri, D., Reiter, R.J., Sewerynek, E., Hara, M., Chen, L., Nistico,<br />

G., 1996. Paraquat toxicity and oxidative damage. Reduction by<br />

melatonin. Biochem. Pharmacol. 51, 1095–1099.<br />

Murray, R.E., Gibson, J.E., 1972. A comparative study <strong>of</strong> <strong>paraquat</strong><br />

<strong>into</strong>xication in rats, guinea pigs and monkeys. Exp. Mol. Pathol.<br />

17, 317–325.<br />

Nemery, B., Smith, L.L., Aldridge, W.N., 1987. Putrescine and 5hydroxytryptamine<br />

accumulation in rat lung slices: cellular localization<br />

and responses to cell-specific lung injury. Toxicol. Appl.<br />

Pharmacol. 91, 107–120.<br />

Onyeama, H.P., Oehme, F.W., 1984. A literature review <strong>of</strong> <strong>paraquat</strong><br />

toxicity. Vet. Hum. Toxicol. 26, 494–502.<br />

R.J. Dinis-Oliveira et al. / Toxicology 227 (2006) 73–85 85<br />

Orzechowski, A., Ostaszewski, P., Wilczak, J., Jank, M., Balasinska,<br />

B., Wareski, P., Fuller, J.J., 2002. Rats with a glucocorticoid<strong>induced</strong><br />

catabolic state show symptoms <strong>of</strong> oxidative stress and<br />

spleen atrophy: <strong>the</strong> effects <strong>of</strong> age and recovery. J. Vet. Med. A<br />

Physiol. Pathol. Clin. Med. 49, 256–263.<br />

Parkinson, C., 1980. The changing pattern <strong>of</strong> <strong>paraquat</strong> poisoning in<br />

man. Histopathology 4, 171–183.<br />

Price, R., 1982. Urinary enzymes, nephrotoxicity and renal disease.<br />

Toxicology 23, 99–134.<br />

Pugin, J., Verghese, G., Widmer, M.C., Matthay, M.A., 1999. The<br />

alveolar space is <strong>the</strong> site <strong>of</strong> intense inflammatory and pr<strong>of</strong>ibrotic<br />

reactions in <strong>the</strong> early phase <strong>of</strong> acute respiratory distress syndrome.<br />

Crit. Care Med. 27, 304–312.<br />

Rannels, D.E., Kameji, R., Pegg, A.E., Rannels, S.R., 1989. Spermidine<br />

uptake by type II pneumocytes: interactions <strong>of</strong> amine<br />

uptake pathways. Am. J. Physiol. Lung Cell Mol. Physiol. 257,<br />

L346–L353.<br />

Rocco, P.R., Souza, A.B., Faffe, D.S., Passaro, C.P., Santos, F.B., Negri,<br />

E.M., Lima, J.G., Contador, R.S., Capelozzi, V.L., Zin, W.A., 2003.<br />

Effect <strong>of</strong> corticosteroid on lung parenchyma remodeling at an early<br />

phase <strong>of</strong> acute lung injury. Am. J. Respir. Crit. Care Med. 168,<br />

677–684.<br />

Schultz, J., Kaminker, K., 1962. Myeloperoxidase <strong>of</strong> <strong>the</strong> leucocyte <strong>of</strong><br />

normal human blood. I. Content and localization. Arch. Biochem.<br />

Biophys. 96, 465–467.<br />

Sharp, C.W., Ottolenghi, A., Posner, H.S., 1972. Correlation <strong>of</strong><br />

<strong>paraquat</strong> toxicity with tissue concentrations and weight loss <strong>of</strong> <strong>the</strong><br />

rat. Toxicol. Appl. Pharmacol. 22, 241–251.<br />

Suzuki, K., Ota, H., Sasagawa, S., Sakatani, T., Fujikura, T., 1983.<br />

Assay method for myeloperoxidase in human polymorphonuclear<br />

leukocytes. Anal. Biochem. 132, 345–352.<br />

Youngman, R.J., Elstner, E.F., 1981. Oxygen species in <strong>paraquat</strong> toxicity:<br />

<strong>the</strong> crypto-OH radical. FEBS Lett. 129, 265–268.


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER V<br />

Full survival <strong>of</strong> <strong>paraquat</strong>-exposed rats after<br />

treatment with sodium salicylate<br />

Reprinted from Free Radical Biology & Medicine 42: 1017-1028<br />

Copyright© (2007) with kind permission from Elsevier Science Inc<br />

177


Part II – Original <strong>research</strong>____________________________________________________<br />

178


Original Contribution<br />

Full survival <strong>of</strong> <strong>paraquat</strong>-exposed rats after treatment with sodium salicylate ☆<br />

R.J. Dinis-Oliveira a,⁎ , C. Sousa a , F. Remião a , J.A. Duarte b , A. Sánchez Navarro c , M.L. Bastos a ,<br />

F. Carvalho a,⁎<br />

a REQUIMTE, Departamento de Toxicologia, Faculdade de Farmácia, Universidade do Porto. Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal<br />

b CIAFEL, Faculdade de Desporto, Universidade do Porto. Rua Dr. Plácido Costa, 91, 4200-450 Porto, Portugal<br />

c Departamento de Farmacia y Tecnología Farmacéutica, Facultad de Farmacia, Universidad de Salamanca. Avda. Campo Charro s/n. 37007, Salamanca, España<br />

Abstract<br />

Received 15 September 2006; revised 21 December 2006; accepted 31 December 2006<br />

Available online 8 January 2007<br />

Over <strong>the</strong> past decades, <strong>the</strong>re have been numerous fatalities resulting from accidental or voluntary ingestion <strong>of</strong> <strong>the</strong> widely used herbicide<br />

<strong>paraquat</strong> dichloride (methyl viologen; PQ). Considering that <strong>the</strong> main target <strong>organ</strong> for PQ toxicity is <strong>the</strong> lung and involves <strong>the</strong> production <strong>of</strong><br />

reactive oxygen and nitrogen species, inflammation, disseminated intravascular coagulation, and activation <strong>of</strong> transcriptional regulatory<br />

<strong>mechanisms</strong>, it may be hypo<strong>the</strong>sized that an antidote against PQ poisonings should counteract all <strong>the</strong>se effects. For this purpose, sodium salicylate<br />

(NaSAL) may constitute an adequate <strong>the</strong>rapeutic drug, due to its ability to modulate inflammatory signaling systems and to prevent oxidative<br />

stress. To test this hypo<strong>the</strong>sis, NaSAL (200 mg/kg ip) was injected in rats 2 h after exposure to a toxic dose <strong>of</strong> PQ (25 mg/kg, ip). NaSAL<br />

treatment caused a significant reduction in PQ-<strong>induced</strong> oxidative stress, platelet activation, and nuclear factor (NF)-κB activation in lung. In<br />

addition, histopathological lesions <strong>induced</strong> by PQ in lung were strongly attenuated and <strong>the</strong> oxidant-<strong>induced</strong> increases <strong>of</strong> glutathione peroxidase<br />

and catalase expression became absent. These effects were associated with a full survival <strong>of</strong> <strong>the</strong> PQ-treated rats (extended for more than 30 days)<br />

in comparison with 100% <strong>of</strong> mortality by Day 6 in animals exposed only to PQ, suggesting that NaSAL constitutes an important and valuable<br />

<strong>the</strong>rapeutic drug to be used against PQ-<strong>induced</strong> toxicity. Indeed, NaSAL constitutes <strong>the</strong> first compound with such degree <strong>of</strong> success (100%<br />

survival).<br />

© 2007 Elsevier Inc. All rights reserved.<br />

Keywords: Oxidative stress; NF-κB; Inflammation; Sodium salicylate; Survival<br />

Introduction<br />

Paraquat dichloride (methyl viologen; PQ) is an effective and<br />

widely used herbicide, which has a proven safety record when<br />

Abbreviations: ARDS, acute respiratory distress syndrome; AUC, area<br />

under curve; CAT, catalase; DNPH, 2,4-dinitrophenylhydrazine; fEMSA,<br />

fluorescence electrophoretic mobility shift assay; GPx, glutathione peroxidase;<br />

H2O2, hydrogen peroxide; HO U , hydroxyl radical; HOCl, hypochlorous acid;<br />

Hyp, hydroxyproline; IκB, inhibitor κB; LM, light microscopy; LPO, lipid<br />

peroxidation; MDA, malondialdehyde; MPO, myeloperoxidase; NaSAL,<br />

sodium salicylate; NF-κB, nuclear factor kappa-B; PMN, polymorphonuclear<br />

leukocytes; PQ, <strong>paraquat</strong>; ROS, reactive oxygen species; SAL, salicylate; TBA,<br />

2-thiobarbituric acid; TBARS, thiobarbituric acid-reactive substances; TCA,<br />

trichloroacetic acid; TEM, transmission electron microscopy; TMB, 3,3′,5,5′tetramethylbenzidine;<br />

TNF-α, tumor necrosis factor alpha.<br />

☆ Portuguese patent pending number 103480.<br />

⁎ Corresponding authors. Fax: +351222003977.<br />

E-mail addresses: ricardinis@ff.up.pt (R.J. Dinis-Oliveira),<br />

felixdc@ff.up.pt (F. Carvalho).<br />

0891-5849/$ - see front matter © 2007 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.freeradbiomed.2006.12.031<br />

Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

www.elsevier.com/locate/freeradbiomed<br />

appropriately applied to eliminate weeds. However, over <strong>the</strong><br />

past decades, <strong>the</strong>re have been numerous fatalities mainly caused<br />

by accidental or voluntary ingestion [1]. The main target <strong>organ</strong><br />

for PQ toxicity is <strong>the</strong> lung as a consequence <strong>of</strong> its accumulation,<br />

against a concentration gradient, through <strong>the</strong> highly developed<br />

polyamine uptake system, and due to its capacity to generate<br />

redox cycle [2–4] (Fig. 1). Death occurs mostly as a<br />

consequence <strong>of</strong> alveolar epi<strong>the</strong>lial cells (type I and II<br />

pneumocytes) and bronchiolar Clara cell disruption, hemorrhage,<br />

edema, hypoxemia, infiltration <strong>of</strong> inflammatory cells <strong>into</strong><br />

<strong>the</strong> interstitial and alveolar spaces, proliferation <strong>of</strong> fibroblasts<br />

and excessive collagen deposition [4], and as a consequence <strong>of</strong> a<br />

disseminated intravascular coagulation [5]. Importantly, platelet<br />

sequestration has been shown to occur in <strong>the</strong> lungs <strong>of</strong> patients<br />

with acute respiratory distress syndrome (ARDS) [6] and<br />

studies have demonstrated platelet effects on membrane<br />

permeability, pulmonary hypertension, activation <strong>of</strong> neutrophils,<br />

endo<strong>the</strong>lial cells, and fibroblasts [7]. Nowadays, no


1018 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Fig. 1. Schematic representation <strong>of</strong> <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. A, cellular diaphorases; SOD, superoxide dismutase or spontaneously; CAT, catalase; GPx,<br />

glutathione peroxidase; Gred, glutathione reductase; PQ 2+ , <strong>paraquat</strong>; PQ .+ , <strong>paraquat</strong> cation radical; HMP, hexose monophosphate pathway; FR, Fenton reaction; HWR,<br />

Haber-Weiss reaction.<br />

antidote or effective treatment for PQ poisoning has been<br />

clinically applied, <strong>the</strong> survival being mainly dependent on <strong>the</strong><br />

amount ingested and <strong>the</strong> time elapsed until <strong>the</strong> patient is<br />

submitted to intensive medical procedures [8].<br />

Despite several studies concerning PQ-<strong>induced</strong> lung<br />

toxicity, few studies focus on <strong>the</strong> transcriptional regulatory<br />

<strong>mechanisms</strong> responsible for PQ toxicity and <strong>the</strong> importance <strong>of</strong><br />

<strong>the</strong> modulation <strong>of</strong> <strong>the</strong>se <strong>mechanisms</strong> in <strong>the</strong> treatment <strong>of</strong> PQ<br />

poisoning. Nuclear factor (NF)-κB has been regarded as a key<br />

element in <strong>the</strong> response <strong>of</strong> cells to inflammatory stimuli. NFκB<br />

activity is attributed to <strong>the</strong> Rel/NF-κB family proteins<br />

forming homo- and heterodimers through a combination <strong>of</strong> <strong>the</strong><br />

subunits p65 (or RelA), p50, p52, c-Rel, and RelB. In most<br />

cells, NF-κB (<strong>the</strong> designation for p50-p65, <strong>the</strong> most frequent<br />

heterodimer) is retained in <strong>the</strong> cytoplasm as an inactive<br />

complex bound to inhibitory proteins [IκB; for revision see<br />

[9]]. LPS, IL-1β, tumor necrosis factor (TNF)-α, UV light,<br />

reactive oxygen species (ROS), and double-stranded RNA are<br />

classical inducers <strong>of</strong> NF-κB. When IκBα is degraded, NF-κB<br />

migrates to <strong>the</strong> nucleus, where it binds to <strong>the</strong> κB sites in <strong>the</strong><br />

promoter region <strong>of</strong> target genes and regulates <strong>the</strong> transcription<br />

<strong>of</strong> proinflammatory enzymes, cytokines, chemokines, apoptosis<br />

inhibitors, cell adhesion molecules, <strong>the</strong> IκBα gene, and<br />

many o<strong>the</strong>rs.<br />

Taking <strong>into</strong> account <strong>the</strong> above-noted rationale, it may be<br />

hypo<strong>the</strong>sized that an antidote against PQ poisonings should<br />

have excellent antioxidant, anti-inflammatory (involving inhibition<br />

<strong>of</strong> NF-κB activation, and antithrombogenic effects. In<br />

that sense, salicylate (SAL) and its derivatives seem to be<br />

adequate candidates for <strong>the</strong> task. SAL is a well-known<br />

scavenger <strong>of</strong> ROS [10] and inhibitor <strong>of</strong> platelet aggregation<br />

[11]. Inhibition <strong>of</strong> <strong>the</strong> NF-κB pathway by SAL has also been<br />

shown by Kopp and collaborators [12] and several o<strong>the</strong>r<br />

subsequent studies by impeding IκB phosphorylation [13,14].<br />

In <strong>the</strong> present work, <strong>the</strong> role <strong>of</strong> oxidative stress, platelet<br />

aggregation, NF-κB activation, and fibrosis in PQ-<strong>induced</strong> lung<br />

toxicity, as well as <strong>the</strong> remarkable healing effects obtained by<br />

<strong>the</strong> administration <strong>of</strong> sodium salicylate (NaSAL), is described.<br />

Importantly, taking <strong>into</strong> account <strong>the</strong> arrival time <strong>of</strong> poisoned<br />

patients to hospital emergencies, NaSAL was administered 2 h<br />

after PQ exposure, conferring more realism to our study. The<br />

obtained results exceeded our best expectations since not only<br />

<strong>the</strong> toxicity was reverted but, most significantly, full survival <strong>of</strong><br />

<strong>the</strong> PQ-<strong>into</strong>xicated rats treated with NaSAL was noted. It may<br />

be postulated that NaSAL is <strong>the</strong> first real PQ antidote described<br />

with such degree <strong>of</strong> success.<br />

Materials and methods<br />

Chemicals and drugs<br />

Paraquat dichloride (1,1′-dimethyl-4,4′-bipyridinium<br />

dichloride), NaSAL (2-hydroxybenzoic acid sodium salt),<br />

3,3′,5,5′-tetramethylbenzidine (TMB), 5-sulfosalicylic acid,<br />

reduced glutathione (GSH), reduced nicotinamide adenine<br />

dinucleotide phosphate (NADPH), hydrogen peroxide (H2O2),<br />

glutathione reductase, 2,4-dinitrophenylhydrazine (DNPH),<br />

trans-4-hydroxy-L-proline, chloramine-T hydrate, and 4-<br />

(dimethylamino)benzaldehyde were all obtained from Sigma<br />

(St. Louis, MO). The saline solution (NaCl 0.9%) and sodium<br />

thiopental were obtained from B. Braun (Lisbon, Portugal). 2-<br />

Thiobarbituric acid (TBA; C 4H 4N 2O 2S), trichloroacetic acid<br />

(TCA; Cl3CCOOH), and sodium hydroxide (NaOH) were<br />

obtained from Merck (Darmstadt, Germany). All <strong>the</strong> reagents<br />

used were <strong>of</strong> analytical grade or from <strong>the</strong> highest available<br />

grade. The following syn<strong>the</strong>tic oligonucleotides, purchased<br />

from Amersham Pharmacia Biotech (Uppsala, Sweden), were<br />

used: 5′-Cy5-GCC TGG GAA AGT CCC CTC AAC T-3′ (NFκB-FW-Cy5),<br />

5′-GCC TGG GAA AGT CCC CTC AAC T-3′<br />

(NF-κB-FW), 5′-AGT TGA GGG GAC TTT CCC AGG C-3′<br />

(NF-κB-R), 5′-CGC TTG ATG ACT CAG CCG GAA-3′ (AP-<br />

1-FW), and 5′-TTC CGG CTG AGT CAT CAA CGC-3′ (AP-<br />

1-R). Cy5 (indodicarbocyanine) is a fluorescence dye attached<br />

at <strong>the</strong> 5′ OH end <strong>of</strong> <strong>the</strong> oligonucleotide. Antibodies against p50<br />

and p65 NF-κB subunits were obtained from Santa Cruz<br />

Biotechnology, Inc.


Animals<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

A total <strong>of</strong> 84 rats were included in <strong>the</strong> study; 44 and 16<br />

animals were used for biochemical and histological studies,<br />

respectively, and <strong>the</strong> remaining (24) for survival rate evaluation.<br />

Male Wistar rats (aged 8 weeks) were obtained from Charles<br />

River S.A. (Barcelona, Spain), with a mean weight <strong>of</strong> 252 ±<br />

25 g. Animals were kept under standard laboratory conditions<br />

(12/12 h light/darkness, 22 ± 2°C room temperature, 50–60%<br />

humidity) for at least 1 week (quarantine) before starting <strong>the</strong><br />

experiments. Animals were allowed access to tap water and rat<br />

chow ad libitum during <strong>the</strong> quarantine period. Animal<br />

experiments were licensed by Portuguese General Directorate<br />

<strong>of</strong> Veterinary Medicine (DGV). Housing and experimental<br />

treatment <strong>of</strong> animals were in accordance with <strong>the</strong> Guide for <strong>the</strong><br />

Care and Use <strong>of</strong> Laboratory Animals from <strong>the</strong> Institute for<br />

Laboratory Animal Research (ILAR 1996). The experiments<br />

complied with <strong>the</strong> current laws <strong>of</strong> Portugal.<br />

Experimental protocol for biochemical and histological studies<br />

The biochemical and histological studies were carried out<br />

with 60 animals randomly, distributed to 10 groups. Each<br />

animal was individually housed during <strong>the</strong> experimental period<br />

in a polypropylene cage with a stainless-steel net at <strong>the</strong> top and<br />

wood chips at <strong>the</strong> screen bottom. Tap water and rat chow were<br />

given ad libitum during <strong>the</strong> entire experiment. Treatments in all<br />

groups were always conducted between 8:00 and 10:00 AM.<br />

Each group was treated as follows (for a schematic view, see<br />

Fig. 2). (i) Control group, n = 6: animals administered with<br />

0.9% NaCl. Animals were administered with one more<br />

administration <strong>of</strong> 0.9% NaCl 2 h later and sacrificed at 24 h<br />

after <strong>the</strong> second injection. (ii) NaSAL group, n = 18: animals<br />

administered with 0.9% NaCl. Animals were treated with one<br />

administration <strong>of</strong> NaSAL (200 mg/kg) 2 h later and sacrificed at<br />

24 h (n = 6, NaSAL 24 h group), 48 h (n = 6, NaSAL 48 h<br />

group), and 96 h (n = 6, NaSAL 96 h group) after <strong>the</strong> second<br />

injection. (iii) PQ group, n = 18: animals <strong>into</strong>xicated with PQ<br />

(25 mg/kg). Animals were administered with one more<br />

administration <strong>of</strong> 0.9% NaCl 2 h later and sacrificed at<br />

24 h (n = 6, PQ 24 h group), 48 h (n = 6, PQ 48 h group),<br />

and 96 h (n = 6, PQ 96 h group) after <strong>the</strong> second injection.<br />

(iv) PQ + NaSAL group, n = 18: animals <strong>into</strong>xicated with PQ<br />

(25 mg/kg). Two hours later, animals were treated with<br />

NaSAL (200 mg/kg) and sacrificed at 24 h (n = 6, PQ +<br />

NaSAL 24 h group), 48 h (n = 6, PQ + NaSAL 48 h group),<br />

and 96 h (n = 6, PQ + NaSAL 96 h group) after <strong>the</strong> second<br />

injection.<br />

The administrations <strong>of</strong> vehicle (0.9% NaCl), PQ, and<br />

NaSAL were all made intraperitoneally (ip) in an injection<br />

volume <strong>of</strong> 1 ml. The experimental dose <strong>of</strong> NaSAL was chosen<br />

in such a way that it covers all <strong>the</strong> desired effects described<br />

above, namely that required to inhibit <strong>the</strong> NF-κB activation in<br />

vivo [13,15]. The PQ administered dose is known to produce<br />

severe lung toxicity and death in rats within a few days<br />

[16,17].<br />

Surgical procedures<br />

Before sacrifice, anes<strong>the</strong>sia was <strong>induced</strong> with sodium<br />

thiopental (60 mg/kg, ip). In four rats <strong>of</strong> each group<br />

(biochemical analysis), lungs were perfused in situ through<br />

<strong>the</strong> inferior vena cava with cold 0.9% NaCl for 3 min at a rate <strong>of</strong><br />

10 ml/min to remove most trapped blood volume. In <strong>the</strong><br />

remaining two animals (structural and ultrastructural analysis),<br />

lung fixation was initiated in situ by perfusion with 2.5%<br />

glutaraldehyde in 0.2 M sodium cacodylate buffer (pH 7.2–7.4)<br />

for 3 min at a rate <strong>of</strong> 10 ml/min. Simultaneous to <strong>the</strong> perfusion<br />

initiation, a cut <strong>of</strong> <strong>the</strong> common iliac artery was done to avoid<br />

cardiovascular volume overload.<br />

Fig. 2. Schematic representation <strong>of</strong> <strong>the</strong> administration protocols for <strong>the</strong> control, sodium salicylate (NaSAL), <strong>paraquat</strong> (PQ), and <strong>paraquat</strong> plus sodium salicylate (PQ +<br />

NaSAL) groups.<br />

1019


1020 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Tissue processing for biochemical analysis<br />

Lungs were removed, cleaned <strong>of</strong> all major cartilaginous<br />

tissues <strong>of</strong> <strong>the</strong> conducting airways, pat-dried with gauze,<br />

weighed [for determination <strong>of</strong> <strong>the</strong> relative lung weight (RLW)<br />

<strong>of</strong> each animal], and processed as follows: (i) Right lungs<br />

(except <strong>the</strong> posterior lobe) were homogenized (1:4 m/v, Ultra-<br />

Turrax homogenizer) in ice-cold 50 mM phosphate buffer with<br />

0.1% (v/v) Triton X-100, pH 7.4. The homogenate was kept on<br />

ice and <strong>the</strong>n centrifuged at 3000g, 4°C, for 10 min. Aliquots <strong>of</strong><br />

<strong>the</strong> resulting supernatants were stored (−80°C) for posterior<br />

quantification <strong>of</strong> myeloperoxidase (MPO), catalase (CAT), and<br />

glutathione peroxidase (GPx) activity, carbonyl groups, hydroxyproline<br />

(Hyp) content, and PQ and protein concentration. The<br />

posterior lobe was homogenized (1:4 m/v, Ultra-Turrax<br />

homogenizer) in TCA 10% and <strong>the</strong>n centrifuged (13,000g,<br />

4°C, for 10 min). Aliquots <strong>of</strong> <strong>the</strong> resulting supernatants were<br />

immediately used to measure <strong>the</strong> degree <strong>of</strong> lipid peroxidation<br />

(LPO). The pellet was used for protein quantification. (ii) Left<br />

lungs were used for preparation <strong>of</strong> nuclear extracts. Briefly, left<br />

lungs were homogenized (Ultra-Turrax homogenizer) in a AC<br />

buffer [(cell lysis buffer), 1 g <strong>of</strong> tissue/3 ml] containing 10 mM<br />

Hepes (pH 7.9), 10 mM KCl, 1.5 mM MgCl2, 0.2% Igepal,<br />

0.5 mM EDTA, 0.1 mM EGTA, 1 mM dithiothreitol (DTT), and<br />

0.25 mM phenylmethylsulfonyl fluoride (PMSF) and incubated<br />

on ice for 15 min. After a brief vortexing, <strong>the</strong> lysates were<br />

centrifuged (850g, 4°C for 10 min). The supernatants<br />

(cytoplasmic extracts) were discharged and <strong>the</strong> pellets were<br />

resuspended (washing step) in 500 μl <strong>of</strong> AC buffer and<br />

incubated for 15 min on ice and <strong>the</strong>n centrifuged (14,000g, 4°C,<br />

for 30 s). The supernatants (cytoplasmic extracts) were<br />

discharged and <strong>the</strong> pellets were resuspended in 500 μl <strong>of</strong>BC<br />

buffer (nuclei lysis buffer) containing 20 mM Hepes, pH 7.9,<br />

420 mM NaCl, 1.5 mM MgCl 2, 2% Igepal, 0.5 mM EDTA, 20%<br />

glycerol, 1 mM DTT, 0.25 mM PMSF, aprotinin (5 μg/ml),<br />

pepsatin (5 μg/ml), leupeptin (5 μg/ml) and incubated on ice for<br />

30 min. After a brief vortexing, <strong>the</strong> lysates were centrifuged<br />

(14,000g, 4°C for 10 min). Supernatants (nuclear extracts) were<br />

collected, divided <strong>into</strong> aliquots, and stored at −80°C for future<br />

NF-κB semiquantification by fluorescent electrophoretic mobility<br />

shift assay (fEMSA). The protein concentration <strong>of</strong> <strong>the</strong><br />

extracts was also determined.<br />

Relative lung weight<br />

The RLW <strong>of</strong> each animal was calculated as a percentage <strong>of</strong><br />

<strong>the</strong> absolute body weight at sacrifice.<br />

Quantification <strong>of</strong> <strong>paraquat</strong> in rat lung<br />

The lung PQ quantification was performed as previously<br />

described [3,16,17].<br />

Tissue processing for structural and ultrastructural analysis<br />

Lung samples were subjected to routine procedures for light<br />

microscopy (LM) and transmission electron microscopy (TEM)<br />

analysis as previously described [16,17]. Histopathological<br />

evidence <strong>of</strong> acute lung damage was semiquantified according to<br />

a previously described procedure [16,17]. For each group, more<br />

than 1000 cells per slide and 100 cells per grid were analyzed in<br />

a blind fashion in order to semiquantify <strong>the</strong> severity and<br />

incidence <strong>of</strong> <strong>the</strong> following parameters: (i) tissue dis<strong>organ</strong>ization,<br />

(ii) inflammatory reaction, (iii) necrotic zones, and (iv)<br />

interstitial fibrosis. The severity <strong>of</strong> tissue dis<strong>organ</strong>ization was<br />

scored according to <strong>the</strong> percentage <strong>of</strong> <strong>the</strong> affected tissue: score 0<br />

= normal structure; score 1 = less than one-third <strong>of</strong> tissue; score<br />

2 = greater than one-third and less than two-thirds; score 3 =<br />

greater than two-thirds <strong>of</strong> tissue. The severity <strong>of</strong> inflammatory<br />

reaction was scored as follows: grade 0 = no cellular infiltration;<br />

grade 1 = mild leukocyte infiltration (1 to 3 cells by visual<br />

field); grade 2 = moderate infiltration (4 to 6 leukocytes by<br />

visual field); and grade 3 = heavy infiltration by neutrophils.<br />

The severity <strong>of</strong> necrosis was scored as follows: grade 0 = no<br />

necrosis; grade 1 = dispersed necrotic foci; grade 2 = confluence<br />

necrotic areas; grade 3 = massive necrosis. The interstitial<br />

fibrosis was scored from 0 (normal lung) to 8 (total fibrosis)<br />

according to <strong>the</strong> following criteria: grade = 0 normal lung; grade<br />

1 = minimal fibrous thickening <strong>of</strong> alveolar or bronchial walls;<br />

grades 2–3 = moderate thickening <strong>of</strong> walls without obvious<br />

damage <strong>of</strong> lung architecture; grades 4–5 = increased fibrosis<br />

with definite damage to lung architecture and formation <strong>of</strong><br />

fibrous bands or small fibrous mass; grades 6–7 = severe<br />

distortion <strong>of</strong> structure and large fibrous areas; “honeycomb<br />

lung” was placed in this category; grade 8 = total fibrotic<br />

obliteration <strong>of</strong> <strong>the</strong> field.<br />

Protein quantification<br />

Protein quantification was performed according to <strong>the</strong><br />

method <strong>of</strong> Lowry et al. [18], using bovine serum albumin as<br />

standard.<br />

Measurement <strong>of</strong> toxicity biomarkers<br />

LPO was evaluated by <strong>the</strong> thiobarbituric acid-reactive<br />

substance (TBARS) methodology [19]. Results are expressed<br />

as nanomole <strong>of</strong> malondialdehyde (MDA) equivalents per<br />

milligram <strong>of</strong> protein using an extinction coefficient (ε) <strong>of</strong><br />

1.56 × 10 5 M −1 cm −1 .<br />

Protein carbonyl groups (ketones and aldehydes) were<br />

determined according to Levine et al. [20]. Results are<br />

expressed as nanomole <strong>of</strong> DNPH incorporated per milligram<br />

<strong>of</strong> protein (ε = 2.2 × 10 4 M −1 cm −1 ).<br />

MPO activity was measured accordingly to <strong>the</strong> method<br />

described before [16,17]. One enzyme unit (U) was defined as<br />

<strong>the</strong> amount <strong>of</strong> enzyme capable to reduce 1 μl<strong>of</strong>H 2O 2/min under<br />

<strong>the</strong> assayed conditions. Results are expressed in enzyme unit per<br />

gram <strong>of</strong> protein (ε = 3.9 × 10 4 M −1 cm −1 ).<br />

CAT activity was measured according to <strong>the</strong> method <strong>of</strong> Aebi<br />

[21]. Results are expressed in enzyme unit per gram <strong>of</strong> protein<br />

(ε = 39.4 M −1 cm −1 ).<br />

GPx activity was measured according to <strong>the</strong> method <strong>of</strong><br />

Flohé and Gunzler [22]. One enzyme unit is equal to


millimole <strong>of</strong> NADPH oxidized per minute per milligram <strong>of</strong><br />

protein. Results are expressed in enzyme unit per gram <strong>of</strong><br />

protein (ε = 6.22 mM −1 cm −1 ).<br />

Hydroxyproline quantification was performed accordingly to<br />

<strong>the</strong> method <strong>of</strong> Reddy and Enwemeka [23], using trans-4hydroxy-L-proline<br />

as standard. Total lung collagen content, as<br />

an index <strong>of</strong> <strong>the</strong> development <strong>of</strong> fibrosis, was calculated with <strong>the</strong><br />

assumption that 12.5% <strong>of</strong> collagen is constituted by hydroxyproline<br />

[24]. Results are expressed as milligram <strong>of</strong> collagen<br />

per gram <strong>of</strong> total protein.<br />

Oligonucleotides and DNA annealing<br />

Oligonucleotides were dissolved in purified water to a final<br />

concentration <strong>of</strong> 0.1 mM prior to use. To generate <strong>the</strong> doublestranded<br />

fluorescent target and <strong>the</strong> equimolar, amounts <strong>of</strong> <strong>the</strong><br />

two complementary single-stranded oligonucleotides (NF-κB-<br />

FW-Cy5 or NF-κB-FW with NF-κB-R and AP-1-FW with AP-<br />

1-R) were mixed in <strong>the</strong> annealing buffer (10 mM Tris-HCl, pH<br />

7.5, 1 mM Na 2EDTA, and 0.5 M NaCl) at final concentrations <strong>of</strong><br />

0.05 mM, heated for 2 min at 95°C (denaturing), incubated for<br />

1 h at 37°C (annealing), and <strong>the</strong>n cooled at 4°C.<br />

Determination <strong>of</strong> transcriptional activation <strong>of</strong> lung nuclear<br />

proteins by fluorescent electrophoretic mobility shift assay<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Fig. 3. Percentage <strong>of</strong> rat surviving in <strong>the</strong> control, <strong>paraquat</strong> (PQ), and <strong>paraquat</strong><br />

plus sodium salicylate (PQ + NaSAL) groups. ccc P < 0.001 versus PQ group.<br />

The NF-κB-binding assay was performed according to a<br />

previously reported method [25]. Nuclear extracts (20 μg <strong>of</strong><br />

protein) were incubated (1 h at 4°C) in a fresh polypropylene<br />

tube with <strong>the</strong> following mixture: 0.5 pmol <strong>of</strong> specific doublestranded<br />

Cy5-labeled for each transcription factor, DNA-binding<br />

buffer [10 mM Hepes (pH 7.9), 0.2 mM EDTA, 50 mM KCl],<br />

2.5 mM DTT, 250 ng <strong>of</strong> poly(dI-dC) · poly(dI-dC), 1% <strong>of</strong> Igepal,<br />

and 10% glycerol. Nine microliters <strong>of</strong> <strong>the</strong> mixture was resolved<br />

by electrophoresis on a 5% nondenaturing polyacrylamide gel at<br />

10°C, 800 V, 50 mA, and 30 W for 3 h in 1X TBE (90 mM Tris<br />

borate, 2 mM EDTA, pH 8.3) using an ALF-Express DNA<br />

sequencer (Amersham Pharmacia Biotech, Uppsala, Sweden).<br />

The temperature was regulated by an external <strong>the</strong>rmostat<br />

ALFexpress II Cooler system (Amersham Pharmacia Biotech).<br />

Specificity <strong>of</strong> <strong>the</strong> DNA–protein complexes was confirmed<br />

by <strong>the</strong> addition <strong>of</strong> a 50-fold excess <strong>of</strong> ei<strong>the</strong>r unlabeled specific<br />

competitor (SC, specific probe without <strong>the</strong> Cy5 label) or<br />

unlabeled nonspecific competitor (UC, which was <strong>the</strong> result <strong>of</strong><br />

<strong>the</strong> annealing <strong>of</strong> AP-1-FW with AP-1-R oligonucleotides). For<br />

supershift assays, antibodies against different NF-κB subunits<br />

p50 and p65 were used. Reactions were identical to gel-shift<br />

reaction conditions except that for supershift assays <strong>the</strong> cells<br />

extracts were preincubated with <strong>the</strong> specific antibody (2 μg) on<br />

ice for 15 min before <strong>the</strong> specific probe was added to <strong>the</strong><br />

mixtures. Signals were analyzed by an ALFwin 1.03 fragment<br />

analyser (Amersham Pharmacia Biotech) and presented as<br />

arbitrary units corresponding to area under curve (AUC).<br />

Experimental protocol for <strong>the</strong> evaluation <strong>of</strong> survival rate<br />

For <strong>the</strong> evaluation <strong>of</strong> survival rate (for a schematic view, see<br />

Fig. 2), 24 animals were randomly divided <strong>into</strong> four groups<br />

(control, NaSAL, PQ, and PQ + NaSAL) <strong>of</strong> six animals each.<br />

Animals were kept under <strong>the</strong> same conditions and treated as<br />

described above. Abnormal findings, including weakness and<br />

dyspnea, were noted and recorded if present. The lethality was<br />

registered every day until Day 30. Rats were weighed daily<br />

during <strong>the</strong> entire study.<br />

Statistical analysis<br />

Results are expressed as mean ± SE. Statistical comparison<br />

between groups was estimated using <strong>the</strong> Kruskal-Wallis<br />

nonparametric method followed by Dunn's test. Comparison<br />

<strong>of</strong> <strong>the</strong> survival curves was performed using <strong>the</strong> logrank test. In all<br />

cases, P values lower than 0.05 were considered as statistically<br />

significant.<br />

Results<br />

Survival rate and macroscopic examination<br />

1021<br />

Rats exposed only to PQ (PQ group) displayed approximately<br />

25 and 100% <strong>of</strong> mortality by <strong>the</strong> second and sixth day,<br />

respectively (Fig. 3). The lethal time (LT)50 for <strong>the</strong> PQ group<br />

was approximately 96 h. Posttreatment <strong>of</strong> PQ-exposed rats with<br />

NaSAL (PQ + NaSAL group) resulted in an enhancement <strong>of</strong> <strong>the</strong><br />

Fig. 4. Rat's body weight for <strong>the</strong> control, <strong>paraquat</strong> (PQ), and <strong>paraquat</strong> plus<br />

sodium salicylate (PQ + NaSAL) groups. ccc P < 0.001 versus PQ group.


1022 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Table 1<br />

Relative lung weight (RLW), PQ levels, and toxicological parameters in <strong>the</strong> control, sodium salicylate (NaSAL), <strong>paraquat</strong> (PQ), and <strong>paraquat</strong> plus sodium salicylate<br />

(PQ + NaSAL) groups<br />

Animal<br />

group<br />

Sacrifice<br />

time<br />

RLW TBARS<br />

(nmol MDA/<br />

mg protein)<br />

Carbonyl<br />

groups<br />

(nmol/mg<br />

protein)<br />

survival time to 100%, 30 days post-PQ injection. These<br />

animals were sacrificed 30 days post-PQ injection and by that<br />

time no signs <strong>of</strong> toxicity were visible. Concerning body weight<br />

(Fig. 4), at <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> study, <strong>the</strong>re were no significant<br />

differences among <strong>the</strong> groups. Significant weight loss occurred<br />

in rats <strong>of</strong> <strong>the</strong> PQ group and in <strong>the</strong> PQ + NaSAL group in <strong>the</strong> first<br />

24 h. Weight stabilized in <strong>the</strong> PQ + NaSAL group between 24<br />

and 48 h and <strong>the</strong>n increased in a similar rate relative to control<br />

and NaSAL groups, during <strong>the</strong> whole study time. No significant<br />

differences were observed in <strong>the</strong> survival time and body weight<br />

between <strong>the</strong> control and <strong>the</strong> NaSAL groups (in order to simplify<br />

only <strong>the</strong> control results are presented).<br />

Diarrhea, piloerection, weight loss, anorexia, adipsia,<br />

hyperpnea, dyspnea, tachycardia, and a red drainage around<br />

<strong>the</strong> mouth, eyes, and nose were present especially in animals<br />

exposed only to PQ during <strong>the</strong> first 48–96 h. During <strong>the</strong> same<br />

experimental period, it was observed that rats belonging to<br />

<strong>the</strong> PQ group only ingested a few milliliters <strong>of</strong> water per day<br />

(8 ± 6 ml) in comparison with 24 ± 10 ml in <strong>the</strong> PQ +<br />

NaSAL and 36 ± 12 ml in <strong>the</strong> control group. Of note, it was<br />

observed that rats <strong>of</strong> <strong>the</strong> PQ + NaSAL 48 h group began to<br />

ingest similar or an even higher (in some cases) amounts <strong>of</strong><br />

water compared to <strong>the</strong> control group.<br />

GPx<br />

(U/mg<br />

protein)<br />

CAT<br />

(U/mg<br />

protein)<br />

Relative lung weight<br />

RLW was assessed as an indication <strong>of</strong> <strong>the</strong> edema degree<br />

(Table 1). No differences were obtained in RLW values among<br />

control, NaSAL, and PQ + NaSAL groups. However, in<br />

comparison to <strong>the</strong>se groups, animals from <strong>the</strong> PQ group showed<br />

a significant RLW increase at 24, 48, and 96 h after PQ exposure<br />

(P < 0.05, P < 0.05, P < 0.01, respectively).<br />

Lung PQ concentrations<br />

MPO<br />

(U/g<br />

protein)<br />

The PQ lung concentrations in <strong>the</strong> PQ and PQ + NaSAL<br />

groups are described in Table 1. Animals posttreated with<br />

NaSAL did not evidence any significant difference concerning to<br />

PQ lung accumulation relatively to <strong>the</strong> PQ-only exposed group.<br />

Structural and ultrastructural analysis<br />

Collagen<br />

(μg/mg<br />

protein)<br />

PQ levels<br />

(μg/mg<br />

protein)<br />

Control 24 h 0.37 ± 0.01 0.20 ± 0.02 1.81 ± 0.08 69.27 ± 0.89 2.52 ± 0.056 22.76 ± 1.90 9.12 ± 1.54<br />

NaSAL 24 h 0.36 ± 0.02 0.20 ± 0.01 1.75 ± 0.08 66.34 ± 1.34 2.50 ± 0.10 18.07 ± 2.03 8.96 ± 0.95<br />

48 h 0.37 ± 0.02 0.19 ± 0.01 1.76 ± 0.07 68.45 ± 2.43 2.50 ± 0.11 19.95 ± 1.11 8.90 ± 2.06<br />

96 h 0.37 ± 0.01 0.21 ± 0.02 1.79 ± 0.05 67.27 ± 2.02 2.46 ± 0.07 19.80 ± 1.86 8.15 ± 1.89<br />

PQ 24 h 0.44 ± 0.02 a,b<br />

0.47 ± 0.02 aaa,bbb 2.28 ± 0.12 a,b<br />

73.96 ± 2.79 a,bb 4.50 ± 0.08 aa,bb 30.20 ± 1.72 a,bb<br />

9.64 ± 0.54 0.134 ± 0.103<br />

48 h 0.49 ± 0.03 a,b<br />

0.40 ± 0.03 aaa,bbb 2.37 ± 0.10 aa,bb 75.16 ± 3.45 a,bb 3.96 ± 0.13 aa,bb 33.75 ± 2.15 aa,bb<br />

9.98 ± 1.09 0.061 ± 0.021<br />

96 h 0.55 ± 0.04 aa,bb 0.42 ± 0.02 aaa,bbb 2.50 ± 0.14 aa,bb 70.65 ± 3.03 3.65 ± 0.08 aa,bb 25.23 ± 2.50 b<br />

11.33 ± 1.12 0.024 ± 0.008<br />

PQ + NaSAL 24 h 0.37 ± 0.03 c<br />

0.32 ± 0.02 a,b,c<br />

1.99 ± 0.09 c<br />

68.96 ± 2.79 c<br />

2.72 ± 0.17 cc<br />

24.68 ± 2.50 c<br />

9.32 ± 1.04 0.137 ± 0.098<br />

48 h 0.38 ± 0.02 c<br />

0.24 ± 0.03 ccc<br />

1.91 ± 0.08 cc<br />

69.32 ± 2.02 c<br />

2.53 ± 0.08 cc<br />

26.45 ± 1.88 c<br />

9.45 ± 1.34 0.057 ± 0.013<br />

96 h 0.36 ± 0.02 cc<br />

0.22 ± 0.02 ccc<br />

1.84 ± 0.06 ccc<br />

70.56 ± 2.88 2.32 ± 0.05 cc<br />

23.26 ± 1.76 9.60 ± 1.73 0.020 ± 0.003<br />

CAT, catalase; MDA, malondialdehyde; GPx, glutathione peroxidase; MPO, myeloperoxidase; NaSAL, sodium salicylate; PQ, <strong>paraquat</strong>; TBARS, thiobarbituric acidreactive<br />

substances.<br />

Values are given as mean ± SE (n = 4). a P < 0.05, aa P < 0.01, and aaa P < 0.001 versus control group, b P < 0.05, bb P < 0.01 and bbb P < 0.001 versus NaSAL group, c P <<br />

0.05, cc P < 0.01 and ccc P < 0.001 versus PQ group.<br />

Major qualitative structural and ultrastructural alterations are<br />

depicted in Fig. 5. Results <strong>of</strong> semiquantitative analysis <strong>of</strong> <strong>the</strong><br />

PQ and PQ + NaSAL groups are presented in Table 2.<br />

Animals from <strong>the</strong> control group presented a normal<br />

pulmonary structure at LM, without evidences <strong>of</strong> alveolar<br />

Fig. 5. Light (A) and electron (B) micrographs from animals <strong>of</strong> control and NaSAL group, respectively, showing a normal pulmonary structure without evidence <strong>of</strong><br />

alveolar collapse, vascular congestion, or cellular infiltrations. Light (C) and electron (D) micrographs from animals injected only with <strong>paraquat</strong> (PQ 24 h group).<br />

Suggestive <strong>of</strong> stasis, it is observed in C an intense vascular congestion with compact and angular erythrocytes (*), and a marked atelectasis (#); in C it is also possible to<br />

identify in <strong>the</strong> alveolar space several phagocytes (pink arrow) and cellular debris (red arrow). In D, beyond endo<strong>the</strong>lial cells with mitochondrial swelling (green<br />

arrows), a capillary filled with angular erythrocytes (*) and numerous activated platelets (*), suggestive <strong>of</strong> an activation <strong>of</strong> <strong>the</strong> blood coagulation system, is also<br />

observed. Light (E) and electron (F) micrographs from animals sacrificed, respectively, 48 and 96 h after PQ exposure. In E necrotic zones (black arrows) <strong>of</strong> alveolar<br />

walls as well as cellular debris (red arrows) and phagocytes (pink arrow) inside alveolar space are visible. It can also be observed several capillaries filled with angular<br />

erythrocytes (*), suggestive <strong>of</strong> vascular stasis. In F are depicted numerous polymorphonuclear within capillaries and in <strong>the</strong> interstitial space (yellow arrows); extended<br />

interstitial areas filled with collagen fibers and fibroblasts (orange arrows) and evidences <strong>of</strong> interstitial edema (#) are also shown; cellular debris (red arrows) and<br />

phagocyte cells (pink arrow) are also observed in <strong>the</strong> narrowed alveolar space as well as several angular erythrocytes within capillaries (*). Light (G) and electron (H)<br />

micrographs from animals <strong>of</strong> PQ + NaSAL 96 h group showing in G necrotic areas (black arrows) and debris (red arrows) and phagocytes cells (pink arrow) within<br />

alveolar space. Noteworthy, is <strong>the</strong> preserved pulmonary structure with several cells containing cytoplasmic inclusions and <strong>the</strong> absence <strong>of</strong> vascular congestion. In H it is<br />

possible to observe cellular debris (red arrows) and a phagocyte within <strong>the</strong> alveolar space (pink arrow); <strong>the</strong> presence <strong>of</strong> edema (#) and collagen fibers (orange arrows) in<br />

<strong>the</strong> interstitial space, and mitochondrial swelling affecting pneumocytes and endo<strong>the</strong>lial cells (green arrows) are also noteworthy.


R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

collapse or cellular infiltrations (Fig. 5A). The TEM<br />

evaluation showed an ordinary alveolar wall, without any<br />

evidences <strong>of</strong> edema or cellular infiltration; <strong>the</strong> pneumocytes<br />

and endo<strong>the</strong>lial cells revealed a preserved ultrastructure<br />

(score 0). Animals from <strong>the</strong> NaSAL group also showed<br />

normal pulmonary structure at 24, 48, and 96 h after NaSAL<br />

administration [score 0 (Fig. 5B)]. On <strong>the</strong> o<strong>the</strong>r hand, PQ<br />

1023<br />

administration <strong>induced</strong> marked alterations compared to <strong>the</strong><br />

control pulmonary pattern, mainly characterized by a diffuse<br />

alveoli collapse with an increased thickness <strong>of</strong> its walls. The<br />

foremost alterations observed in <strong>the</strong> animals <strong>of</strong> <strong>the</strong> PQ 24 h<br />

group include intense vascular congestion with activated<br />

platelets and numerous polymorphonuclear cells inside <strong>the</strong><br />

capillaries, apparently adherent to endo<strong>the</strong>lial cells (Figs. 5C


1024 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Table 2<br />

Semiquantitative analysis <strong>of</strong> <strong>the</strong> morphological injury parameters <strong>of</strong> <strong>the</strong> <strong>paraquat</strong> (PQ) and <strong>paraquat</strong> plus sodium salicylate (PQ + NaSAL) groups<br />

PQ PQ + NaSAL<br />

24 h 48 h 96 h 24 h 48 h 96 h<br />

Tissue dis<strong>organ</strong>ization 2.1 ± 0.23 a,b<br />

2.3 ± 0.21 a,b<br />

2.6 ± 0.16 a,b<br />

2.0 ± 0.21 a,b<br />

1.8 ± 0.25 a,b<br />

1.4 ± 0.16 a,b,c<br />

Inflammatory reaction 2.0 ± 0.26 a,b<br />

2.2 ± 0.20 a,b<br />

2.1 ± 0.18 a,b<br />

2.1 ± 0.23 a,b<br />

2.3 ± 0.26 a,b<br />

1.9 ± 0.20 a,b<br />

Necrotic zones 1.2 ± 0.20 1.7 ± 0.33 a,b<br />

2.6 ± 0.16 a<br />

1.3 ± 0.21 1.6 ± 0.22 a,b<br />

1.5 ± 0.17 a,b,c<br />

Interstitial fibrosis 0.2 ± 0.13 0.7 ± 0.21 2.9 ± 0.3 a,b<br />

0.5 ± 0.22 1.1 ± 0.23 1.8 ± 0.27 a,b,c<br />

Values are given as mean ± SE (n = 2). a P < 0.05 versus control group, b P < 0.05 versus NaSAL group, c P < 0.05 versus PQ group.<br />

and 5D). Animals also revealed an interstitial edema,<br />

indicated by <strong>the</strong> existence <strong>of</strong> intercellular vacuolization<br />

areas that were characterized by a minor density ultrastructure<br />

at TEM. The majority <strong>of</strong> pneumocytes and endo<strong>the</strong>lial<br />

cells showed, at least, one ultrastructural abnormality, with<br />

mitochondrial swelling as <strong>the</strong> most frequent alteration (Fig.<br />

5D). These histopathological alterations became more<br />

exuberant at 48 and 96 h after PQ exposure, with <strong>the</strong><br />

necrotic areas, <strong>the</strong> fibroblast activation, and extent <strong>of</strong><br />

interstitial areas occupied by collagen fibers being particularly<br />

notorious (Figs. 5E and 5F). Noteworthy, in <strong>the</strong> PQ +<br />

NaSAL 24, 48, and 96 h groups, compared to <strong>the</strong> PQ-only<br />

exposed animals, <strong>the</strong> occurrence <strong>of</strong> <strong>the</strong> above referred<br />

alterations was drastically attenuated, particularly <strong>the</strong> vascular<br />

congestion (Figs. 5G and 5H).<br />

NF-κB<br />

fEMSA was performed to study <strong>the</strong> effects <strong>of</strong> PQ<br />

exposure in <strong>the</strong> activation <strong>of</strong> rat lung NF-κB. As shown in<br />

Fig. 6, PQ <strong>induced</strong> a significant and time-dependent<br />

activation <strong>of</strong> NF-κB in rat lungs (Lanes 2–4) compared to<br />

control (Lane 1) and NaSAL groups (Lanes 8–10).<br />

Noteworthy was also <strong>the</strong> significant reduction <strong>of</strong> NF-κB<br />

lung activation in <strong>the</strong> NaSAL 24 h group relative to control.<br />

Concerning <strong>the</strong> PQ + NaSAL 24, 48, and 96 groups (Lanes<br />

11, 12, and 13, respectively), NaSAL treatment resulted in a<br />

significant reduction <strong>of</strong> PQ-<strong>induced</strong> NF-κB activation, <strong>the</strong><br />

AUC <strong>of</strong> band 1 being near to that <strong>of</strong> <strong>the</strong> control group. The<br />

specificity <strong>of</strong> <strong>the</strong> DNA–protein complex was confirmed in<br />

<strong>the</strong> PQ 96 h group by maintenance <strong>of</strong> <strong>the</strong> bands in <strong>the</strong><br />

competition experiment with a 50-fold molar excess <strong>of</strong> <strong>the</strong><br />

UC (Lane 6) and by its disappearance in <strong>the</strong> competition<br />

experiment with a 50-fold molar excess <strong>of</strong> <strong>the</strong> SC (Lane 7).<br />

Supershift analysis <strong>of</strong> <strong>the</strong> lung samples <strong>of</strong> rats from <strong>the</strong> PQ<br />

96 h group, using p50 and p65 antibodies, confirmed <strong>the</strong><br />

specificity <strong>of</strong> NF-κB bands (data not shown). The antibody<br />

against subunit p65 was able to shift DNA/protein<br />

interaction present in band 1. The antibody against subunit<br />

p50 shifted band 2 and <strong>induced</strong> a partial decrease in band<br />

1. Taken toge<strong>the</strong>r, <strong>the</strong>se results indicated that p50/p65<br />

heterodimers and p50/p50 homodimers corresponded to<br />

complexes/bands 1 and 2, respectively. It is interesting to<br />

observe that most <strong>of</strong> <strong>the</strong> transcriptional activity altered by<br />

PQ administration is mediated through <strong>the</strong> complex p50/<br />

p65, which is considered <strong>the</strong> most frequent active form <strong>of</strong><br />

NF-κB.<br />

Glutathione peroxidase and catalase activities<br />

PQ produced a significant increase in <strong>the</strong> GPx and CAT<br />

activities in lung tissue when compared with control and<br />

NaSAL groups (Table 1). Treatment with NaSAL attenuated <strong>the</strong><br />

PQ-<strong>induced</strong> increase <strong>of</strong> <strong>the</strong>se enzymes activities.<br />

Lipid peroxidation and carbonyl group content<br />

As shown in Table 1, animals from <strong>the</strong> PQ group exhibited a<br />

significant rise <strong>of</strong> MDA concentration in lung at 24, 48, and<br />

96 h post-PQ exposure (P < 0.001), compared with animals<br />

from control and NaSAL groups. However, as can be observed<br />

in animals from <strong>the</strong> PQ + NaSAL group, NaSAL led to a<br />

significant reduction in lung MDA equivalents compared to<br />

animals from <strong>the</strong> PQ group in every sampled time. Of note, <strong>the</strong><br />

increase <strong>of</strong> LPO in <strong>the</strong> PQ + NaSAL 24 h group was also<br />

significantly higher in comparison to control or <strong>the</strong> NaSAL 24 h<br />

group (P < 0.05). Analogous results were also obtained for<br />

carbonyl group levels (Table 1). In accordance with <strong>the</strong> results<br />

<strong>of</strong> LPO, NaSAL treatment prevented <strong>the</strong> increase <strong>of</strong> protein<br />

carbonylation by PQ in all groups.<br />

Myeloperoxidase activity<br />

Aliquots <strong>of</strong> rat lung samples were assayed for <strong>the</strong> activity <strong>of</strong><br />

MPO as an index <strong>of</strong> lung invasion by neutrophils. As shown in<br />

Table 1, lung MPO activities <strong>of</strong> <strong>the</strong> PQ 24 and 48 h groups were<br />

significantly higher than in rats from control or <strong>the</strong> NaSAL<br />

group. Concerning <strong>the</strong> PQ 96 h group, a significant difference<br />

was only achieved in comparison to <strong>the</strong> NaSAL 96 h group (P <<br />

0.05). Posttreatment with NaSAL prevented <strong>the</strong> increase <strong>of</strong><br />

MPO activity as a consequence <strong>of</strong> PQ exposure in all groups.<br />

Hydroxyproline lung content<br />

Despite <strong>the</strong> enhanced fibrotic changes observed in <strong>the</strong> lung<br />

histology <strong>of</strong> PQ-exposed animals relative to control or NaSAL<br />

groups, <strong>the</strong> biochemical measurement <strong>of</strong> Hyp was not sensitive<br />

enough to detect it in <strong>the</strong>se earlier stages. Hyp contents <strong>of</strong> <strong>the</strong><br />

lung tissue were comparable among control, PQ, PQ + NaSAL,<br />

and NaSAL groups (Table 1).<br />

Discussion<br />

The results obtained in <strong>the</strong> present study clearly show that<br />

NaSAL confers a potent protection against PQ-<strong>induced</strong> lung


R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

Fig. 6. Time course <strong>of</strong> NF-κB activation <strong>induced</strong> by PQ in lungs. fEMSA gel view and binding activities, presented as arbitrary units corresponding to area under curve<br />

(AUC), are shown. Nuclear extracts from <strong>the</strong> different groups were prepared and subjected to fEMSA as described under Materials and methods. Lane 1, control group<br />

(C); Lanes 2, 3 and 4, <strong>paraquat</strong> (PQ) 24, 48, and 96 h groups, respectively; Lane 5, blank (B); Lane 6, competition experiment with a 50-fold molar excess <strong>of</strong> a<br />

nonspecific competitor (UC) compared to specific probe (SP); Lane 7, competition experiment with a 50-fold molar excess <strong>of</strong> a specific competitor (SC, unlabeled<br />

specific probe) compared to SP. Lanes 8, 9, and 10, sodium salicylate (NaSAL) 24, 48, and 96 h groups, respectively; Lanes 11, 12, and 13, PQ + NaSAL 24, 48, and<br />

96 h groups, respectively; The positions <strong>of</strong> specific NF-κB/DNA-binding complexes (bands 1–3) are indicated. NS Band represents a nonspecific binding. The<br />

localization <strong>of</strong> <strong>the</strong> free probe (FP) is also indicated. The AUC results are given as mean ± SE (n = 4). a P < 0.05, aa P < 0.01 and aaa P < 0.001 versus control group, b P <<br />

0.05, bb P < 0.01, and bbb P < 0.001 versus NaSAL group, c P < 0.05, cc P < 0.01, and ccc P < 0.001 versus PQ group.<br />

toxicity. It is shown, for <strong>the</strong> first time, that <strong>the</strong> administration <strong>of</strong><br />

NaSAL (200 mg/kg ip), 2 h after PQ (25 mg/kg ip) exposure,<br />

results in a remarkable decrease <strong>of</strong> PQ-<strong>induced</strong> lung toxicity in<br />

Wistar rats. The prevention <strong>of</strong> PQ-<strong>induced</strong> lung toxicity by<br />

NaSAL was evidenced by a significant remission <strong>of</strong> several<br />

biochemical and histopathological biomarkers <strong>of</strong> toxicity, and<br />

resulted in full survival <strong>of</strong> <strong>the</strong> <strong>into</strong>xicated animals. Importantly,<br />

NaSAL was administered 2 h after PQ exposure, conferring<br />

more realism to our study since, in <strong>the</strong> majority <strong>of</strong> <strong>the</strong> PQ<br />

<strong>into</strong>xications, medical assistance is only possible a few hours<br />

after PQ <strong>into</strong>xication.<br />

As expected, LPO and carbonyl group levels increased<br />

significantly in <strong>the</strong> lung <strong>of</strong> rats exposed to PQ compared to <strong>the</strong><br />

control group. These oxidative stress-related alterations, which<br />

are in agreement with our previous reports [16,17], were<br />

attenuated by NaSAL administration. A plausible justification<br />

for this protection conferred by NaSAL is its potent scavenging<br />

effect on hydroxyl radical (HO U ). Among ROS, HO U is thought<br />

to be <strong>the</strong> most damaging species and <strong>the</strong> mainly responsible for<br />

protein oxidation and LPO [26]. Indeed, <strong>the</strong> major hydroxylation<br />

products <strong>of</strong> HO U attack on SAL, 2,3-dihydroxybenzoic<br />

acid (2,3-DHBA) and 2,5-dihydroxybenzoic acid (2,5-DHBA),<br />

have been used as sensitive markers for <strong>the</strong> measurement <strong>of</strong><br />

1025<br />

HO U formation in vivo [27]. Thus, <strong>the</strong> efficacy <strong>of</strong> NaSAL in<br />

preventing PQ-<strong>induced</strong> lung injury may also be due to its ability<br />

to inactivate <strong>the</strong> HO U radical. In addition, local biotransformation<br />

<strong>of</strong> NaSAL could play an important role in its ability to<br />

markedly attenuate lung oxidative injury due to PQ exposure,<br />

since 2,3-DHBA is a potent iron chelator [28] and iron<br />

availability is required for HO U generation via Fenton reaction.<br />

In addition to <strong>the</strong> increase <strong>of</strong> oxidative damage markers, it<br />

was also observed that NF-κB expression underwent a<br />

significant and sustained increase in <strong>the</strong> lung as a consequence<br />

<strong>of</strong> PQ exposure. The induction <strong>of</strong> NF-κB expression was time<br />

dependent, which seems to point to a continuous inflammatory<br />

process that ultimately may cause severe lung damage and<br />

death. Indeed, a large oral dose <strong>of</strong> PQ (>30 mg/kg in humans)<br />

rapidly leads to multi<strong>organ</strong> failure, with lung damage consisting<br />

<strong>of</strong> disruption <strong>of</strong> alveolar epi<strong>the</strong>lial cells (type I and II<br />

pneumocytes) and bronchiolar Clara cells (CC), hemorrhage,<br />

edema, hypoxemia, and infiltration <strong>of</strong> inflammatory cells <strong>into</strong><br />

<strong>the</strong> interstitial and alveolar spaces [29]. To our knowledge this is<br />

<strong>the</strong> first study describing <strong>the</strong> association <strong>of</strong> a strong NF-κB<br />

activation in <strong>the</strong> lung along with <strong>the</strong> aggravation <strong>of</strong> <strong>the</strong> health<br />

status <strong>of</strong> rats <strong>into</strong>xicated with PQ. NaSAL strongly suppressed<br />

<strong>the</strong> PQ-<strong>induced</strong> lung NF-κB activation, which probably


1026 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

contributes to <strong>the</strong> observed healing effect mediated by this drug.<br />

NF-κB induction by PQ may be correlated with <strong>the</strong> antioxidant<br />

enzyme levels, as previously shown by Zhou and co-workers<br />

[30], who exposed skeletal muscle cells to PQ-<strong>induced</strong><br />

oxidative stress and demonstrated that NF-κB mediates <strong>the</strong><br />

induction <strong>of</strong> GPx and CAT. NaSAL, given 2 h after PQ,<br />

attenuated <strong>the</strong> increase <strong>of</strong> CAT and GPx activities near to <strong>the</strong><br />

levels <strong>of</strong> control and NaSAL groups, which is in accordance<br />

with <strong>the</strong> reported correlation. Accordingly, <strong>the</strong> same rational<br />

could also be applied to explain <strong>the</strong> increase <strong>of</strong> GPx and CAT<br />

activities by lung resident cells in PQ-only exposed animals.<br />

However, it is important to stress <strong>the</strong> positive time change<br />

correlation between <strong>the</strong> antioxidant enzymes and <strong>the</strong> MPO<br />

activities, observed in <strong>the</strong>se animals. Despite suggesting an<br />

association linking <strong>the</strong> infiltration <strong>of</strong> inflammatory cells and <strong>the</strong><br />

enhancement <strong>of</strong> antioxidant enzymes, with our experimental<br />

design it was not possible to establish a direct cause–effect<br />

relationship and thus distinguish <strong>the</strong> individual contribution <strong>of</strong><br />

<strong>the</strong> polymorphonuclear leukocytes (PMN) antioxidant enzymes<br />

arsenal from <strong>the</strong> CAT and GPx overexpression by lung cells as a<br />

consequence <strong>of</strong> PQ exposure. Our biochemical results showed<br />

that MPO activity decreased in <strong>the</strong> lungs <strong>of</strong> <strong>the</strong> PQ + NaSAL<br />

groups compared to PQ-only exposed groups. Since MPO is<br />

located within <strong>the</strong> primary azurophil granules <strong>of</strong> PMN, its<br />

activity indirectly reflects PMN infiltration through <strong>the</strong> <strong>organ</strong>s<br />

[31] during <strong>the</strong> inflammatory reaction. It could be argued that<br />

<strong>the</strong> observed NaSAL protective effects against PQ-<strong>induced</strong> lung<br />

toxicity may be <strong>the</strong> result <strong>of</strong> less infiltration by inflammatory<br />

cells. Histopathological studies confirmed <strong>the</strong> widespread<br />

neutrophil infiltration especially in <strong>the</strong> lungs <strong>of</strong> PQ-only<br />

exposed animals. Macrophage infiltration and several NK<br />

cells were also identified in <strong>the</strong> interstitial space (Figs. 5C and<br />

5D). It has been shown that <strong>the</strong> adhesion <strong>of</strong> circulating PMN to<br />

vascular endo<strong>the</strong>lium is crucial to <strong>the</strong>ir transendo<strong>the</strong>lial<br />

migration [32]. The expression <strong>of</strong> adhesion molecules on<br />

endo<strong>the</strong>lial cells is regulated by NF-κB [33]. As a consequence<br />

<strong>of</strong> IκB phosphorylation inhibition, NaSAL has proven to inhibit<br />

transendo<strong>the</strong>lial migration <strong>of</strong> neutrophils [14,33]. However,<br />

histopathological results revealed only some amelioration <strong>of</strong><br />

PMN lung infiltration, in contrast to <strong>the</strong> drastic decrease <strong>of</strong><br />

MPO activity observed in rats <strong>of</strong> PQ + NaSAL groups in<br />

comparison to <strong>the</strong> PQ groups. According to <strong>the</strong> literature, this<br />

apparent discrepancy between <strong>the</strong> biochemical and <strong>the</strong><br />

histopathological results could be at least partially explained<br />

by a lower generation <strong>of</strong> <strong>the</strong> most powerful oxidant produced by<br />

human neutrophils, HOCl, as a consequence <strong>of</strong> MPO inhibition<br />

by NaSAL [34]. Thus, it is reasonable to consider that <strong>the</strong><br />

NaSAL overall protection could also be <strong>the</strong> consequence <strong>of</strong><br />

MPO inhibition. In <strong>the</strong> present study, PQ also caused lung<br />

edema, an effect observed as an increase <strong>of</strong> RLW and by<br />

histopathological analysis, which confirms <strong>the</strong> great contribution<br />

<strong>of</strong> inflammation to <strong>the</strong> toxic effects mediated by this<br />

herbicide. Exuberance <strong>of</strong> interstitial edema was drastically<br />

attenuated in PQ + NASAL animals.<br />

There are two distinct phases in <strong>the</strong> development <strong>of</strong><br />

pulmonary fibrosis (PF) resulting from PQ exposure. The first<br />

is a destructive phase in which <strong>the</strong> alveolar type I and type II and<br />

Clara cells are destroyed within 1–3 days after PQ poisoning,<br />

resulting in alveolitis [4]. This stage provides a basis for an<br />

extensive PF observed in <strong>the</strong> second phase, since <strong>the</strong> cells<br />

involved in <strong>the</strong> alveolitis, e.g., macrophages, lymphocytes, and<br />

neutrophils play a key role in producing <strong>the</strong> factors that regulate<br />

<strong>the</strong> proliferation, chemotactism, and secretory activity <strong>of</strong><br />

fibroblasts and consequently <strong>the</strong> extent <strong>of</strong> <strong>the</strong> interstitial and<br />

intraalveolar fibrosis [35,36]. Alveolar macrophages secrete<br />

fibrogenic factors such as transforming growth factor (TGF)-β<br />

[37] and gene expression <strong>of</strong> TGF-β is enhanced in <strong>the</strong> lungs<br />

after PQ exposure [38]. Activated macrophages in inflamed<br />

lungs in response to PQ exposure also syn<strong>the</strong>size increased<br />

amounts <strong>of</strong> several o<strong>the</strong>r cytokines, including IL-1α, IL-1β, IL-<br />

6, platelet-derived growth factor (PDGF)-A, TGF-α, insulin<br />

like growth factor, TNF-α, and monocyte chemoattractant<br />

protein (MCP)-1 that mediate an enhanced fibroproliferative<br />

response [39]. Accordingly, an increase <strong>of</strong> local TNF-α<br />

production was observed in bleomycin-<strong>induced</strong> PF [40] and<br />

Ishida et al. [38] found an up-regulation <strong>of</strong> TNF-α mRNA<br />

expression in lungs <strong>of</strong> PQ-exposed mice with a concomitant<br />

increase in leukocyte infiltration. All <strong>the</strong>se events represent<br />

possible causes for unremitting lung fibrosis as a consequence<br />

<strong>of</strong> PQ <strong>into</strong>xication as well as potential targets for <strong>the</strong>rapeutic<br />

intervention. On its own, NaSAL has been shown to blunt <strong>the</strong><br />

increase in TNF-α mRNA and to reduce <strong>the</strong> serum TNF-α<br />

protein level <strong>of</strong> mice [41]. The fact that <strong>the</strong>se events have been<br />

observed in <strong>the</strong> early inflammatory phase suggests that <strong>the</strong><br />

effect <strong>of</strong> NaSAL may be related to tissue preservation; that is, a<br />

decreased extent <strong>of</strong> lung lesion at <strong>the</strong> initial phase in NaSALtreated<br />

animals would imply a reduced fibrotic process at later<br />

stages. This hypo<strong>the</strong>sis does not exclude a direct effect <strong>of</strong><br />

NaSAL on preventing fibrogenesis, since it has been reported<br />

that acetylsalicylic acid inhibits collagen syn<strong>the</strong>sis by fibroblast<br />

proliferation inhibition in vitro [42]. Taking <strong>into</strong> account <strong>the</strong><br />

capacity <strong>of</strong> NF-κB to activate <strong>the</strong> transcription <strong>of</strong> specific genes<br />

and classes <strong>of</strong> genes encoding for various proinflammatory and<br />

fibrogenic cytokines, it is reasonable to consider that <strong>the</strong><br />

inhibition <strong>of</strong> <strong>the</strong> NF-κB activation plays an important role in <strong>the</strong><br />

NaSAL protective effect against PQ-<strong>induced</strong> lung toxicity. Our<br />

results also confirm earlier studies regarding <strong>the</strong> fibrogenic<br />

effects <strong>of</strong> PQ. Collagen lung deposition was observed in a timedependent<br />

manner by histology 48 and 96 h after PQ exposure.<br />

Despite <strong>the</strong> enhanced fibrotic changes observed in <strong>the</strong> lung<br />

histology <strong>of</strong> PQ-exposed animals relative to control or NaSAL<br />

groups <strong>the</strong> biochemical measurement <strong>of</strong> Hyp was not sensitive<br />

enough to detect it in <strong>the</strong>se earlier stages. Hyp contents <strong>of</strong> <strong>the</strong><br />

lung tissue were comparable among control, PQ, PQ + NaSAL,<br />

and NaSAL groups. Previous studies, where Hyp lung content<br />

was determined as a measure <strong>of</strong> fibrosis, showed that an<br />

increase in its levels was observed only after 14 days <strong>of</strong> PQ<br />

exposure [43,44]. The lack <strong>of</strong> utility for Hyp measurement at <strong>the</strong><br />

first days after PQ exposure was also confirmed in our study.<br />

The histological results showed an attenuation <strong>of</strong> collagen<br />

deposition in <strong>the</strong> PQ + NaSAL 96 h group, which support a<br />

beneficial effect <strong>of</strong> NaSAL in preventing PF.<br />

Ano<strong>the</strong>r point <strong>of</strong> interest resulting from this study comes<br />

from <strong>the</strong> histopathological analysis. Rats from <strong>the</strong> PQ group


R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

evidenced an intense vascular congestion, <strong>the</strong> pulmonary<br />

capillaries being filled with angular erythrocytes and numerous<br />

activated platelets, suggestive <strong>of</strong> an activation <strong>of</strong> <strong>the</strong><br />

blood coagulation system. These findings are in accordance<br />

with <strong>the</strong> disseminated intravascular coagulation described for<br />

PQ [5]. One <strong>of</strong> <strong>the</strong> main results <strong>of</strong> NaSAL inclusion in <strong>the</strong><br />

<strong>the</strong>rapeutic <strong>of</strong> PQ-poisoned rats was <strong>the</strong> reduction <strong>of</strong> vascular<br />

congestion without signs <strong>of</strong> platelet activation, which may be,<br />

at least partially, explained by <strong>the</strong> antithrombogenic effect <strong>of</strong><br />

NaSAL [11]. It is well known that PQ causes ARDS, which<br />

may be <strong>the</strong> result <strong>of</strong> blood coagulation and vascular stasis<br />

leading to an enhanced pulmonary dead-space fraction [45].<br />

Since blood coagulation evidence in <strong>the</strong> NaSAL-treated<br />

animals was attenuated, it is legitimate to speculate that <strong>the</strong><br />

ratio <strong>of</strong> ventilation/perfusion was not severely disturbed as in<br />

<strong>the</strong> PQ group, justifying <strong>the</strong> survival rate observed in <strong>the</strong>se<br />

animals.<br />

We have recently demonstrated that <strong>the</strong> induction <strong>of</strong> de<br />

novo syn<strong>the</strong>sis <strong>of</strong> membrane P-glycoprotein (P-gp) by<br />

dexamethasone confers a strong protection against PQ-<strong>induced</strong><br />

lung toxicity by increasing its efflux from <strong>the</strong> lung [17].<br />

According to <strong>the</strong> literature, we could not disregard this<br />

possibility occurring as well as for NaSAL, since aspirin has<br />

proven to enhance expression <strong>of</strong> P-gp and thus to induce<br />

multidrug resistance [46,47]. We could nei<strong>the</strong>r ignore <strong>the</strong><br />

possibility <strong>of</strong> NaSAL to increase <strong>the</strong> PQ elimination from<br />

pneumocytes, by raising its export by <strong>the</strong> polyamine export<br />

transporter [48]. However, our results did not point to <strong>the</strong><br />

protection occurring by decreasing PQ lung accumulation,<br />

since no statistical difference was observed between <strong>the</strong> PQ<br />

and <strong>the</strong> PQ + NaSAL groups.<br />

In view <strong>of</strong> our results, it is plausible to conclude that a<br />

<strong>the</strong>rapy with NaSAL, starting as soon as possible after PQ<br />

<strong>into</strong>xication, may constitute a promising treatment for PQ<br />

poisonings. One may consider that <strong>the</strong> dose <strong>of</strong> NaSAL used in<br />

this study is quite high. According to <strong>the</strong> literature, <strong>the</strong><br />

pathophysiologic changes attributable to high doses <strong>of</strong> NaSAL<br />

result in various clinical manifestations depending on <strong>the</strong><br />

amount ingested; in humans, an oral dose <strong>of</strong> 200 mg/kg yields<br />

approximately a serum concentration <strong>of</strong> 500–770 mg/L [49–<br />

51]. These serum levels originate signals <strong>of</strong> mild side effects<br />

such as nausea, vomiting, tinnitus, hyperventilation, and<br />

respiratory alkalosis. In <strong>the</strong> particular case <strong>of</strong> life-threatening<br />

PQ poisonings, <strong>the</strong> risk/benefit ratio will most probably flip to<br />

<strong>the</strong> beneficial effects <strong>of</strong> NaSAL, although this still needs to be<br />

carefully confirmed in clinical trials. In addition, NaSAL<br />

proved to protect lungs, which was confirmed by a remission<br />

<strong>of</strong> practically all toxicological parameters that were changed in<br />

<strong>the</strong> lung <strong>of</strong> PQ-challenged rats, through an effective inhibition<br />

<strong>of</strong> proinflammatory factors, scavenging <strong>of</strong> ROS, inhibition <strong>of</strong><br />

MPO, and inhibition <strong>of</strong> platelet aggregation. This protection<br />

achieved full survival <strong>of</strong> <strong>the</strong> PQ-exposed animals.<br />

In our opinion this <strong>the</strong>rapeutic approach has <strong>the</strong> potential to<br />

be applied in humans, though fur<strong>the</strong>r preclinical studies are<br />

needed, particularly those aimed to explain in more detail <strong>the</strong><br />

mode <strong>of</strong> action <strong>of</strong> this interesting molecule in <strong>the</strong> protection<br />

against PQ-<strong>induced</strong> lung damage.<br />

Acknowledgment<br />

Ricardo Dinis-Oliveira acknowledges FCT for his Ph.D.<br />

grant (SFRH/BD/13707/2003).<br />

References<br />

1027<br />

[1] Bismuth, C., Hall, A. H. (Eds.). Paraquat poisoning: <strong>mechanisms</strong>,<br />

prevention, treatment. Dekker, New York; 1995.<br />

[2] Dinis-Oliveira, R. J.; Remião, F.; Duarte, J. A.; Sánchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. Paraquat exposure as an etiological factor <strong>of</strong><br />

Parkinson's disease. Neurotoxicology 27:1110–1122; 2006.<br />

[3] Dinis-Oliveira, R. J.; Valle, M. J.; Bastos, M. L.; Carvalho, F.; Sánchez-<br />

Navarro, A. Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung. Influence <strong>of</strong><br />

sodium depletion. Xenobiotica 36:724–737; 2006.<br />

[4] Lewis, C. P.; Nemery, B. Pathophysiology and biochemical <strong>mechanisms</strong> <strong>of</strong><br />

<strong>the</strong> pulmonary toxicty <strong>of</strong> <strong>paraquat</strong>. In: Bismuth, C., Hall, A. H. (Eds.).<br />

(Eds.), Paraquat poisoning: <strong>mechanisms</strong>, prevention, treatment, vol. 10.<br />

Dekker, NewYork, pp. 107–140; 1995.<br />

[5] Harrison, L. C.; Dortimer, A. C.; Murphy, K. J. Fatalities due to <strong>the</strong> weedkiller<br />

<strong>paraquat</strong>. Med. J. Aust. 2:774–777; 1972.<br />

[6] Schneider, R. C.; Zapol, W. M.; Carvalho, A. C. Platelet consumption and<br />

sequestration in severe acute respiratory failure. Am. Rev. Respir. Dis.<br />

122:445–451; 1980.<br />

[7] Heffner, J. E.; Sahn, S. A.; Repine, J. E. The role <strong>of</strong> platelets in <strong>the</strong> adult<br />

respiratory distress syndrome. Culprits or bystanders? Am. Rev. Respir.<br />

Dis.; 1987.<br />

[8] Dinis-Oliveira, R. J.; Sarmento, A.; Reis, P.; Amaro, A.; Remião, F.;<br />

Bastos, M. L.; Carvalho, F. Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival<br />

case following intake <strong>of</strong> a potential lethal dose. Pediatr. Emerg. Care<br />

22:537–540; 2006.<br />

[9] Whiteside, S. T.; Israel, A. I kappa B proteins: structure, function and<br />

regulation. Semin. Cancer Biol. 8:75–82; 1997.<br />

[10] Jaarsveld, H.; van Kuyl, J. M.; Zyl, G. F.; van Barnard, H. C. Salicylate in<br />

<strong>the</strong> perfusate during ischemia/reperfusion prevented mitochondrial injury.<br />

Res. Commun. Mol. Pathol. Pharmacol. 86:287–295; 1994.<br />

[11] Roncaglioni, M. C.; Reyers, I.; Cerletti, C.; Donati, M. B.; de Gaetano, G.<br />

Moderate anticoagulation by salicylate prevents thrombosis without<br />

bleeding complications. An experimental study in rats. Biochem.<br />

Pharmacol. 37:4743–4745; 1988.<br />

[12] Kopp, E.; Ghosh, S. Inhibition <strong>of</strong> NF-kappa B by sodium salicylate and<br />

aspirin. Science 265:956–959; 1994.<br />

[13] Huang, C. J.; Tsai, P. S.; Lu, Y. T.; Cheng, C. R.; Stevens, B. R.; Skimming,<br />

J. W.; Pan, W. H. NF-kappaB involvement in <strong>the</strong> induction <strong>of</strong> high affinity<br />

CAT-2 in lipopolysaccharide-stimulated rat lungs. Acta Anaes<strong>the</strong>siol.<br />

Scand. 48:992–1002; 2004.<br />

[14] Pierce, J. W.; Read, M. A.; Ding, H.; Luscinskas, F. W.; Collins, T.<br />

Salicylates inhibit I kappa B-alpha phosphorylation, endo<strong>the</strong>lial-leukocyte<br />

adhesion molecule expression, and neutrophil transmigration. J. Immunol.<br />

156:3961–3969; 1996.<br />

[15] Yang, C. H.; Tsai, P. S.; Lee, J. J.; Huang, C. H.; Huang, C. J. NF-kappaB<br />

inhibitors stabilize <strong>the</strong> mRNA <strong>of</strong> high-affinity type-2 cationic amino acid<br />

transporter in LPS-stimulated rat liver. Acta Anaes<strong>the</strong>siol. Scand.<br />

49:468–476; 2005.<br />

[16] Dinis-Oliveira, R. J.; Duarte, J. A.; Remiao, F.; Sanchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. Single high dose dexamethasone treatment<br />

decreases <strong>the</strong> pathological effects and increases <strong>the</strong> survival rat <strong>of</strong><br />

<strong>paraquat</strong>-<strong>into</strong>xicated rats. Toxicology 227:73–85; 2006.<br />

[17] Dinis-Oliveira, R. J.; Remião, F.; Duarte, J. A.; Sanchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. P-glycoprotein induction: an antidotal pathway<br />

for <strong>paraquat</strong>-<strong>induced</strong> lung toxicity. Free Radic. Biol. Med. 41:1213–1224;<br />

2006.<br />

[18] Lowry, O. H. N.; Rosebrough, N. J.; Farr, A. L.; Randall, R. J. Protein<br />

measurement with Folin phenol reagent. J. Biol. Chem. 193:265–275;<br />

1951.<br />

[19] Buege, J. A.; Aust, S. D. Microsomal lipid peroxidation. Methods<br />

Enzymol. 52:302–310; 1978.


1028 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 42 (2007) 1017–1028<br />

[20] Levine, R. L.; Williams, J. A.; Stadtman, E. R.; Shacter, E. Carbonyl assay<br />

for determination <strong>of</strong> oxidatively modified proteins. Methods Enzymol.<br />

233:346–357; 1994.<br />

[21] Aebi, H. Catalase in vitro. Methods Enzymol. 105:121–126; 1984.<br />

[22] Flohé, L.; Gunzler, W. A. Assays <strong>of</strong> glutathione peroxidase. Methods<br />

Enzymol. 105:114–121; 1984.<br />

[23] Reddy, G. K.; Enwemeka, C. S. A simplified method for <strong>the</strong> analysis <strong>of</strong><br />

hydroxyproline in biological tissues. Clin. Biochem. 29:225–229; 1996.<br />

[24] Edwards, C. A.; O'Brien Jr., W. D. Modified assay for determination <strong>of</strong><br />

hydroxyproline in a tissue hydrolyzate. Clin. Chim. Acta 104:161–167;<br />

1980.<br />

[25] Ruscher, K.; Reuter, M.; Kupper, D.; Trendelenburg, G.; Dirnagl, U.;<br />

Meisel, A. A fluorescence based non-radioactive electrophoretic mobility<br />

shift assay. J. Biotechnol. 78:163–170; 2000.<br />

[26] Kaur, H.; Halliwell, B. Detection <strong>of</strong> hydroxyl radicals by aromatic<br />

hydroxylation. Methods Enzymol. 233:67–82; 1994.<br />

[27] Coudray, C.; Talla, M.; Martin, S.; Fatome, M.; Favier, A. Highperformance<br />

liquid chromatography-electrochemical determination <strong>of</strong><br />

salicylate hydroxylation products as an in vivo marker <strong>of</strong> oxidative<br />

stress. Anal. Biochem. 227:101–111; 1995.<br />

[28] Graziano, J. H.; Grady, R. W.; Cerami, A. The identification <strong>of</strong> 2, 3dihydroxybenzoic<br />

acid as a potentially useful iron-chelating drug. J.<br />

Pharmacol. Exp. Ther. 190:570–575; 1974.<br />

[29] Onyeama, H. P.; Oehme, F. W. A literature review <strong>of</strong> <strong>paraquat</strong> toxicity. Vet.<br />

Hum. Toxicol. 26:494–502; 1984.<br />

[30] Zhou, L. Z.; Johnson, A. P.; Rando, T. A. NF kappa B and AP-1 mediate<br />

transcriptional responses to oxidative stress in skeletal muscle cells. Free<br />

Radic. Biol. Med. 31:1405–1416; 2001.<br />

[31] Schultz, J.; Kaminker, K. Myeloperoxidase <strong>of</strong> <strong>the</strong> leucocyte <strong>of</strong> normal<br />

human blood. I. Content and localization. Arch. Biochem. Biophys.<br />

96:465–467; 1962.<br />

[32] Springer, T. A. Traffic signals on endo<strong>the</strong>lium for lymphocyte recirculation<br />

and leukocyte emigration. Annu. Rev. Physiol. 57:827–872; 1995.<br />

[33] Collins, T.; Read, M. A.; Neish, A. S.; Whitley, M. Z.; Thanos, D.; Maniatis,<br />

T. Transcriptional regulation <strong>of</strong> endo<strong>the</strong>lial cell adhesion molecules: NFkappa<br />

B and cytokine-inducible enhancers. FASEB J. 9:899–909; 1995.<br />

[34] Kettle, A. J.; Winterbourn, C. C. Mechanism <strong>of</strong> inhibition <strong>of</strong><br />

myeloperoxidase by anti-inflammatory drugs. Biochem. Pharmacol.<br />

41:1485–1492; 1991.<br />

[35] Smith, R. E.; Strieter, R. M.; Phan, S. H.; Lukacs, N. W.; Huffnagle, G. B.;<br />

Wilke, C. A.; Burdick, M. D.; Lincoln, P.; Evan<strong>of</strong>f, H.; Kunkel, S. L.<br />

Production and function <strong>of</strong> murine macrophage inflammatory protein-1<br />

alpha in bleomycin-<strong>induced</strong> lung injury. J. Immunol. 153:4704–4712;<br />

1994.<br />

[36] Strausz, J.; Muller-Quernheim, J.; Steppling, H.; Ferlinz, R. Oxygen<br />

radical production by alveolar inflammatory cells in idiopathic pulmonary<br />

fibrosis. Am. Rev. Respir. Dis. 141:124–128; 1990.<br />

[37] Zhang, Y.; Lee, T. C.; Guillemin, B.; Yu, M. C.; Rom, W. N. Enhanced IL-<br />

1 beta and tumor necrosis factor-alpha release and messenger RNA<br />

expression in macrophages from idiopathic pulmonary fibrosis or after<br />

asbestos exposure. J. Immunol. 150:4188–4196; 1993.<br />

[38] Ishida, Y.; Takayasu, T.; Kimura, A.; Hayashi, T.; Kakimoto, N.;<br />

Miyashita, T.; Kondo, T. Gene expression <strong>of</strong> cytokines and growth<br />

factors in <strong>the</strong> lungs after <strong>paraquat</strong> administration in mice. Leg. Med.<br />

(Tokyo) 8:102–109; 2006.<br />

[39] Khalil, N.; O'Connor, R. N. Cytokine regulation <strong>of</strong> pulmonary fibrosis. In:<br />

Phan, S. M., Thrall, R. S. (Eds.), Pulmonary Fibrosis, vol. 80. Dekker, New<br />

York, pp. 627–646; 1995.<br />

[40] Piguet, P. F.; Collart, M. A.; Grau, G. E.; Kapanci, Y.; Vassalli, P. Tumor<br />

necrosis factor/cachectin plays a key role in bleomycin-<strong>induced</strong><br />

pneumopathy and fibrosis. J. Exp. Med. 170:655–663; 1989.<br />

[41] Ramesh, G.; Reeves, W. B. Salicylate reduces cisplatin nephrotoxicity<br />

by inhibition <strong>of</strong> tumor necrosis factor-alpha. Kidney Int. 65:490–499;<br />

2004.<br />

[42] Surazynski, A.; Palka, J.; Wolczynski, S. Acetylsalicylic acid-dependent<br />

inhibition <strong>of</strong> collagen biosyn<strong>the</strong>sis and beta1-integrin signaling in cultured<br />

fibroblasts. Med. Sci. Monit. 10:BR175–BR179; 2004.<br />

[43] Kuttan, R.; Lafranconi, M.; Sipes, I. G.; Meezan, E.; Brendel, K. Effect <strong>of</strong><br />

<strong>paraquat</strong> treatment on prolyl hydroxylase activity and collagen syn<strong>the</strong>sis <strong>of</strong><br />

rat lung and kidney. Res. Commun. Chem. Pathol. Pharmacol. 25:<br />

257–268; 1979.<br />

[44] Fukushima, T.; Tanaka, K.; Lim, H.; Moriyama, M. Changes in <strong>the</strong> fatty<br />

acid composition and hydroxyproline content in rat lung in relation to<br />

collagen syn<strong>the</strong>sis after <strong>paraquat</strong> administration. Fukushima J. Med. Sci.<br />

49:33–43; 2003.<br />

[45] Rocco, P. R.; Souza, A. B.; Faffe, D. S.; Passaro, C. P.; Santos, F. B.; Negri,<br />

E. M.; Lima, J. G.; Contador, R. S.; Capelozzi, V. L.; Zin, W. A. Effect <strong>of</strong><br />

corticosteroid on lung parenchyma remodeling at an early phase <strong>of</strong> acute<br />

lung injury. Am. J. Respir. Crit. Care Med. 168:677–684; 2003.<br />

[46] Flescher, E.; Rotem, R.; Kwon, P.; Azare, J.; Jaspers, I.; Cohen, D. Aspirin<br />

enhances multidrug resistance gene 1 expression in human Molt-4 T<br />

lymphoma cells. Anticancer Res. 20:4441–4444; 2000.<br />

[47] Rotem, R.; Tzivony, Y.; Flescher, E. Contrasting effects <strong>of</strong> aspirin on<br />

prostate cancer cells: suppression <strong>of</strong> proliferation and induction <strong>of</strong> drug<br />

resistance. Prostate 42:172–180; 2000.<br />

[48] Hughes, A.; Smith, N. I.; Wallace, H. M. Polyamines reverse non-steroidal<br />

anti-inflammatory drug-<strong>induced</strong> toxicity in human colorectal cancer cells.<br />

Biochem. J. 374:481–488; 2003.<br />

[49] Dargan, P. I.; Wallace, C. I.; Jones, A. L. An evidence based flowchart to<br />

guide <strong>the</strong> management <strong>of</strong> acute salicylate (aspirin) overdose. Emerg. Med.<br />

J. 19:206–209; 2002.<br />

[50] Proudfoot, A. T. Toxicity <strong>of</strong> salicylates. Am. J. Med. 75:99–103; 1983.<br />

[51] Yip, L.; Dart, R. C.; Gabow, P. A. Concepts and controversies in salicylate<br />

toxicity. Emerg. Med. Clin. North Am. 12:351–364; 1994.


____________________________________________________Part II – Original <strong>research</strong><br />

CHAPTER VI<br />

Effects <strong>of</strong> sodium salicylate in <strong>the</strong> <strong>paraquat</strong>-<strong>induced</strong><br />

apoptotic events in rat lungs<br />

Reprinted from Free Radical Biology & Medicine 43: 48-61<br />

Copyright© (2007) with kind permission from Elsevier Science Inc<br />

191


Part II – Original <strong>research</strong>____________________________________________________<br />

192


Original Contribution<br />

Sodium salicylate prevents <strong>paraquat</strong>-<strong>induced</strong> apoptosis in <strong>the</strong> rat lung<br />

R.J. Dinis-Oliveira a,⁎ , C. Sousa a , F. Remião a , J.A. Duarte b , R. Ferreira b , A. Sánchez Navarro c ,<br />

M.L. Bastos a , F. Carvalho a,⁎<br />

a REQUIMTE, Departamento de Toxicologia, Faculdade de Farmácia, Universidade do Porto, Rua Aníbal Cunha, 164, 4099-030 Porto, Portugal<br />

b CIAFEL, Faculdade de Desporto, Universidade do Porto, Rua Dr. Plácido Costa, 91, 4200-450 Porto, Portugal<br />

c Departamento de Farmacia y Tecnología Farmacéutica, Facultad de Farmacia, Universidad de Salamanca, Avda. Campo Charro s/n. 37007, Salamanca, España<br />

Abstract<br />

Received 2 November 2006; revised 11 March 2007; accepted 13 March 2007<br />

Available online 24 March 2007<br />

The nonselective contact herbicide, <strong>paraquat</strong> (PQ), is a strong pneumotoxicant, especially due to its accumulation in <strong>the</strong> lung through a<br />

polyamine uptake system and to its capacity to induce redox cycling, leading to oxidative stress-related damage. In <strong>the</strong> present study, we aimed to<br />

investigate <strong>the</strong> occurrence <strong>of</strong> apoptotic events in <strong>the</strong> lungs <strong>of</strong> male Wistar rats, 24, 48, and 96 h after PQ exposure (25 mg/kg ip) as well as <strong>the</strong><br />

putative healing effects provided by sodium salicylate [(NaSAL), 200 mg/kg ip] when administered 2 h after PQ. PQ exposure resulted in marked<br />

lung apoptosis, in a time-dependent manner, characterized by <strong>the</strong> “ladder-like” pattern <strong>of</strong> DNA observed through electrophoresis and by <strong>the</strong><br />

presence <strong>of</strong> terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate nick end-labeling (TUNEL)-positive cells (TPC) as revealed<br />

by immunohistochemistry. The two main caspase cascades (<strong>the</strong> extrinsic receptor-mediated and <strong>the</strong> intrinsic mitochondria-mediated) and <strong>the</strong><br />

expressions <strong>of</strong> p53 and activator protein-1 (AP-1) were also evaluated, to obtain an insight <strong>into</strong> apoptotic cellular signaling. PQ-exposed rats<br />

suffered a time-dependent increase <strong>of</strong> caspase-3 and caspase-8 and a decrease <strong>of</strong> caspase-1 activities in lungs compared to <strong>the</strong> control group. A<br />

marked mitochondrial dysfunction evidenced by cytochrome c (Cyt c) release was also observed as a consequence <strong>of</strong> PQ exposure. In addition,<br />

fluorescence electrophoretic mobility shift assay (fEMSA) revealed a transcriptional induction <strong>of</strong> <strong>the</strong> p53 and AP-1 transcription factors in a timedependent<br />

manner as a consequence <strong>of</strong> PQ exposure. NaSAL treatment resulted in <strong>the</strong> remission <strong>of</strong> <strong>the</strong> observed apoptotic signaling and<br />

consequently <strong>of</strong> lung apoptosis. Taken toge<strong>the</strong>r, <strong>the</strong> present results showed that PQ activates several events involved in <strong>the</strong> apoptotic pathways,<br />

which might contribute to its lung toxicodynamics. NaSAL, a recently implemented antidote for PQ <strong>into</strong>xications, proved to protect lungs from PQ<strong>induced</strong><br />

apoptosis.<br />

© 2007 Elsevier Inc. All rights reserved.<br />

Keywords: Paraquat; Lung toxicity; Sodium salicylate; Apoptosis; Rats<br />

Free Radical Biology & Medicine 43 (2007) 48–61<br />

www.elsevier.com/locate/freeradbiomed<br />

Abbreviations: AC buffer, cell lysis buffer; AP-1, activator protein-1; Apaf-1, apoptosis-activating factor-1; BC buffer, nuclei lysis buffer; CAT, catalase; Chaps, 3-<br />

[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate); Cyt c, cytochrome c; DTT, dithiothreitol; dUTP, fluorescein-labeled deoxyuridinetriphosphate; ECL,<br />

enhanced chemiluminescence; FADD, Fas-associating protein with death domain; fEMSA, fluorescence electrophoretic mobility shift assay; GPx, glutathione<br />

peroxidase; HO· , hydroxyl radical; LPO, lipid peroxidation; LPS, lipopolysaccharide; MPO, myeloperoxidase; NaSAL, sodium salicylate; NF-κB, nuclear factor<br />

kappa-B; PBS, phosphate-buffered saline; PCD, programmed cell death; PMSF, phenylmethylsulfonyl fluoride; PQ, <strong>paraquat</strong>; RNase A, ribonuclease A; ROS,<br />

reactive oxygen species; SC, specific competitor; SDS, sodium dodecyl sulfate; TdT, terminal deoxynucleotidyl transferase; TBARS, thiobarbituric acid-reactive<br />

substances; TNF-α, tumor necrosis factor alpha; TPC, TUNEL-positive cells; TRAIL, TNF-related apoptosis-inducing ligand; TUNEL, deoxynucleotidyl transferasemediated<br />

deoxyuridine triphosphate nick end-labeling; UC, unspecific competitor.<br />

⁎ Corresponding authors. Fax: +351 222003977.<br />

E-mail addresses: ricardinis@ff.up.pt (R.J. Dinis-Oliveira), felixdc@ff.up.pt (F. Carvalho).<br />

0891-5849/$ - see front matter © 2007 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.freeradbiomed.2007.03.014


Introduction<br />

The main target <strong>organ</strong> for <strong>the</strong> toxicity elicited by <strong>the</strong><br />

herbicide <strong>paraquat</strong> dichloride (methyl viologen; PQ) is <strong>the</strong> lung.<br />

The <strong>mechanisms</strong> subjacent to this <strong>organ</strong> specificity are<br />

postulated to be associated with <strong>the</strong> selective accumulation <strong>of</strong><br />

PQ in <strong>the</strong> lung, followed by a sustained redox-cycling effect,<br />

leading to oxidative stress-related cell death and inflammation<br />

[1,2].<br />

Despite numerous studies concerning PQ-<strong>induced</strong> toxicity,<br />

few <strong>of</strong> <strong>the</strong>m focus on <strong>the</strong> apoptotic and on <strong>the</strong> transcriptional<br />

regulatory <strong>mechanisms</strong> as potential contributory factors for PQ<br />

toxicity and <strong>the</strong> importance <strong>of</strong> modulating <strong>the</strong>se <strong>mechanisms</strong> in<br />

<strong>the</strong> treatment <strong>of</strong> PQ poisonings. PQ-<strong>induced</strong> apoptosis was first<br />

demonstrated in a murine myeloid cell line, mouse 32D cells<br />

[3]. Subsequently, <strong>the</strong> involvement <strong>of</strong> reactive oxygen species<br />

(ROS) in <strong>the</strong> occurrence <strong>of</strong> PQ-<strong>induced</strong> apoptosis was reported,<br />

ei<strong>the</strong>r using <strong>the</strong> in vivo model <strong>of</strong> <strong>the</strong> intrahippocampal injection<br />

<strong>of</strong> PQ [4] or using <strong>the</strong> in vitro models <strong>of</strong> differentiated human<br />

neuroblastoma cells (SHSY-5Y) [5] and human lung epi<strong>the</strong>lial<br />

cells [6]. Notwithstanding <strong>the</strong> lack <strong>of</strong> in vivo studies about <strong>the</strong><br />

putative apoptotic effects <strong>of</strong> PQ in <strong>the</strong> lung, it is known that<br />

alveolar epi<strong>the</strong>lium undergoes apoptosis in normal tissue<br />

remodeling as well as in pathological conditions [7,8]. Our<br />

hypo<strong>the</strong>sis is that <strong>the</strong> selective uptake <strong>of</strong> PQ by type I and II<br />

pneumocytes, and Clara cells [9] may lead to a subsequent<br />

induction <strong>of</strong> apoptosis, which could make <strong>the</strong> lung cells unable<br />

to restore normal tissue architecture and function, leading to<br />

irreversible damage.<br />

Apoptosis or programmed cell death (PCD) is an essential<br />

process <strong>of</strong> cell death during embryonic and postnatal tissue<br />

remodeling as well as in several pathological conditions [10].<br />

Morphologically, apoptosis is characterized by reduction <strong>of</strong> cell<br />

volume, membrane blebbing, chromatin condensation, nuclear<br />

fragmentation, and apoptotic cell body formation [11]. The<br />

signaling pathways leading to apoptosis are implemented by a<br />

death machinery signaling system whose executionary arm is a<br />

family <strong>of</strong> cysteine proteases, designated caspases (for cysteine<br />

aspartic acid-specific proteases) [12–14]. Caspases exist<br />

normally as inactive precursors (procaspases) in <strong>the</strong> cytosolic<br />

fraction <strong>of</strong> <strong>the</strong> cells. They are cleaved proteolytically at specific<br />

amino acid sequences <strong>into</strong> low molecular weight units (20–<br />

23 kDa), when <strong>the</strong> cell undergoes apoptosis, to form <strong>the</strong> active<br />

enzyme. Two main caspase cascades, <strong>the</strong> extrinsic receptor<br />

mediated and intrinsic mitochondria mediated, have been<br />

delineated in mammalian cells [12–14]. The extrinsic pathway<br />

for <strong>the</strong> activation <strong>of</strong> apoptosis involves <strong>the</strong> stimulation <strong>of</strong> death<br />

receptors expressed at <strong>the</strong> cell surface, leading to clustering and<br />

formation <strong>of</strong> a death-inducing signaling complex system, which<br />

includes <strong>the</strong> adapter protein FADD (Fas-associated death<br />

domain) and <strong>the</strong> initiator caspase-8 [15]. Both TNF-related<br />

apoptosis-inducing ligand (TRAIL) and tumor necrosis factor<br />

(TNF)-α are known to bind to <strong>the</strong>ir cell surface receptors,<br />

leading to caspase-8 activation. Caspase-8 is a major enzyme<br />

activating downstream <strong>the</strong> effector caspase-3 [16]. The intrinsic<br />

pathway involves <strong>the</strong> release <strong>of</strong> cytochrome c (Cyt c) from<br />

mitochondria to <strong>the</strong> cytosolic fraction <strong>of</strong> <strong>the</strong> cells at an early<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

phase <strong>of</strong> apoptosis [12–14,17]. Cyt c [also known as apoptosisactivating<br />

factor-2 (Apaf-2)], toge<strong>the</strong>r with some cytosolic<br />

proteins (i.e., Apaf-1) in <strong>the</strong> presence <strong>of</strong> dATP [13], recruits and<br />

activates <strong>the</strong> conversion <strong>of</strong> <strong>the</strong> latent apoptosis-promoting<br />

procaspase-9 to its active form, which <strong>the</strong>n activates caspase-3<br />

[13,17,18]. Thus, both death receptor and mitochondria pathways<br />

converge at <strong>the</strong> level <strong>of</strong> caspase-3 activation. Caspase-3 is<br />

considered to be <strong>the</strong> central and final apoptotic effector enzyme<br />

responsible for many <strong>of</strong> <strong>the</strong> biological, morphological, and<br />

structural features <strong>of</strong> apoptosis [12]. Active effector caspases<br />

mediate <strong>the</strong> cleavage <strong>of</strong> apoptosis regulators, <strong>the</strong> cleavage <strong>of</strong><br />

housekeeping proteins, and DNA fragmentation, resulting in<br />

morphological features <strong>of</strong> apoptosis [13,14].<br />

The first caspase to be identified, caspase-1 [interleukin (IL)-<br />

1β-converting enzyme], is not a component <strong>of</strong> cell death<br />

machinery, but it indirectly influences <strong>the</strong> rates <strong>of</strong> apoptosis<br />

through cleavage <strong>of</strong> IL-1β to its 17-kDa mature form [19].<br />

Rowe et al. [20] hypo<strong>the</strong>sized that caspase-1 might regulate<br />

apoptosis <strong>of</strong> neutrophils. These investigators studied <strong>the</strong>se<br />

processes in caspase-1-deficient mice compared with wild-type<br />

controls. The results provided evidence for a proapoptotic role<br />

<strong>of</strong> caspase-1 in <strong>the</strong> lipopolysaccharide (LPS)-unstimulated neutrophils,<br />

what is reversed in LPS-treated neutrophils by <strong>the</strong><br />

antiapoptotic effects mediated by IL-1β cleavage. In addition,<br />

using a model <strong>of</strong> LPS-mediated lung injury, <strong>the</strong>y found that<br />

caspase-1-deficient mice show a prolonged inflammatory response<br />

[20].<br />

O<strong>the</strong>r actors in <strong>the</strong> apoptotic story are <strong>the</strong> transcription factor<br />

activator protein-1 (AP-1) and <strong>the</strong> tumor suppressor protein<br />

p53. AP-1 is <strong>the</strong> designation <strong>of</strong> <strong>the</strong> transcriptional complex<br />

composed <strong>of</strong> dimers (homodimeric and heterodimeric) <strong>of</strong> proteins<br />

<strong>of</strong> <strong>the</strong> fos (c-fos, fos-B, fra1, and fra2) and jun oncogene<br />

families (c-jun, jun-B, and jun-D) [21]. Much <strong>of</strong> what is known<br />

about <strong>the</strong> biological function <strong>of</strong> AP-1 relates to its prominent<br />

roles in cell proliferation, differentiation, and transformation<br />

and in <strong>the</strong> induction <strong>of</strong> apoptosis [22,23]. Its activation occurs in<br />

response to a number <strong>of</strong> diverse stimuli, including oxidative or<br />

cellular stress, ultraviolet irradiation, DNA damage, antigen<br />

binding by T or B lymphocytes, and exposure to proinflammatory<br />

cytokines (e.g., TNF-α, transforming growth factor-β, and<br />

γ-interferon), overlapping in several instances with <strong>the</strong> target<br />

genes <strong>of</strong> NF-κB [23]. In addition to increased subunit syn<strong>the</strong>sis,<br />

oxidative stress induces AP-1-mediated transcription by<br />

enhancing DNA-binding activity as well [24]. With <strong>the</strong><br />

exception <strong>of</strong> preexisting c-jun homodimers, induction <strong>of</strong> AP-1<br />

relies predominantly on novel syn<strong>the</strong>sis <strong>of</strong> its DNA-binding<br />

subunits [25]. The p53 tumor suppressor protein is a 53-kDa<br />

transcription factor constitutively expressed at low levels in<br />

most cells and tissues [26]. It is presumably <strong>the</strong> most intensively<br />

studied factor <strong>of</strong> programmed cell death, since its overexpression<br />

induces apoptosis [26,27]. Several lines <strong>of</strong> evidence,<br />

supporting a key role for p53 in <strong>the</strong> control <strong>of</strong> cell cycle <strong>of</strong> a<br />

range <strong>of</strong> cell types by controlling <strong>the</strong> progression through G1phase,<br />

have arisen [26,27].<br />

Recently, our group demonstrated that sodium salicylate<br />

(NaSAL) constitutes an important and valuable antidote to be<br />

used against PQ-<strong>induced</strong> toxicity, leading to full survival <strong>of</strong> PQ-<br />

49


50 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

exposed rats. The antidotal effect <strong>of</strong> NaSAL was mainly a<br />

consequence <strong>of</strong> <strong>the</strong> effective inhibition <strong>of</strong> <strong>the</strong> proinflammatory<br />

factor, nuclear factor (NF)-κB, scavenging <strong>of</strong> ROS, inhibition<br />

<strong>of</strong> myeloperoxidase (MPO), and inhibition <strong>of</strong> platelet aggregation<br />

[28]. Therefore, <strong>the</strong> same approach was followed here,<br />

primarily to determine <strong>the</strong> ability <strong>of</strong> PQ to induce apoptotic<br />

events in <strong>the</strong> lungs <strong>of</strong> Wistar rats and secondly to evaluate if <strong>the</strong><br />

treatment with NaSAL has also beneficial effects at this level.<br />

Apoptosis was assessed by <strong>the</strong> “ladder-like” pattern <strong>of</strong> DNA<br />

and by <strong>the</strong> terminal deoxynucleotidyl transferase-mediated<br />

deoxyuridine triphosphate nick end-labeling (TUNEL) assay.<br />

Concerning apoptotic cell signaling, <strong>the</strong> two main caspase<br />

cascades were studied through <strong>the</strong> measurement <strong>of</strong> <strong>the</strong> cytosolic<br />

Cyt c concentrations and <strong>of</strong> <strong>the</strong> enzymatic activities <strong>of</strong> caspases-1,<br />

-8, and -3. The expressions <strong>of</strong> p53 and AP-1 were also<br />

evaluated. Overall, this study should lead to a better understanding<br />

<strong>of</strong> <strong>the</strong> underlying adverse pathways activated by PQ in<br />

<strong>the</strong> respiratory tract and consequently to provide new tools to<br />

prevent PQ-<strong>induced</strong> lung toxicity.<br />

Materials and methods<br />

Chemicals and drugs<br />

PQ (1,1′-dimethyl-4,4′-bipyridinium dichloride), NaSAL (2hydroxybenzoic<br />

acid sodium salt), N-acetyl-Trp-Glu-His-Aspp-nitroanilide<br />

(colorimetric substrate for caspase-1), N-acetyl-<br />

Asp-Glu-Val-Asp p-nitroanilide (colorimetric substrate for<br />

caspase-3), N-acetyl-Ile-Glu-Thr-Asp-p-nitroanilide (colorimetric<br />

substrate for caspase-8), Chaps (3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate),<br />

ribonuclease A<br />

(RNase A), sodium dodecyl sulfate (SDS), proteinase K from<br />

Tritirachium album, phenol solution (equilibrated with<br />

10 mM Tris-HCl, pH 8.0, 1 mM EDTA), chlor<strong>of</strong>orm, Mayer's<br />

hematoxylin solution, eosin Y, and <strong>the</strong> alkaline phosphatase<br />

substrate solution (Fast Red TR/Napthol AS-MX) were<br />

obtained from Sigma (St. Louis, MO). Anti-cytochrome c<br />

antibody (556433) was obtained from BD Pharmigen. Mouse<br />

monoclonal anti-α-tubulin (clone DM1A) was purchased from<br />

Lab Vision Corporation (Fremont, CA). The enhanced chemiluminescence<br />

(ECL)-Plus reagent and <strong>the</strong> entire Western blot<br />

reagents were purchased from Amersham Biosciences (Lisbon,<br />

Portugal). The saline solution (NaCl 0.9%) and sodium<br />

thiopental were obtained from B. Braun (Lisbon, Portugal).<br />

The following syn<strong>the</strong>tic oligonucleotides, purchased from<br />

Amersham Pharmacia Biotech (Uppsala, Sweden), were used:<br />

5′-Cy5-TAC AGA ACA TGT CTA AGC ATG CTG GGG-3′<br />

(p53-FW-Cy5), 5′-TAC AGA ACA TGT CTA AGC ATG CTG<br />

GGG-3′ (p53-FW), 5′-CCC CAG CAT GCT TAG ACA TGT<br />

TCT GTA-3′ (p53-FW-R), 5′-Cy5-CGC TTG ATG ACT CAG<br />

CCG GAA-3′ (AP-1-FW-Cy5), 5′-CGC TTG ATG ACT CAG<br />

CCG GAA-3′ (AP-1-FW), 5′-TTC CGG CTG AGT CAT CAA<br />

CGC-3′ (AP-1-R), 5′-GCC TGG GAA AGT CCC CTC AAC<br />

T-3′ (NF-κB-FW), and 5′-AGT TGA GGG GAC TTT CCC<br />

AGG C-3′ (NF-κB-R). Cy5 (indodicarbocyanine) is a fluorescence<br />

dye attached to <strong>the</strong> 5′ OH end <strong>of</strong> <strong>the</strong> oligonucleotide. All<br />

<strong>the</strong> reagents used were <strong>of</strong> analytical grade.<br />

Animals<br />

A total <strong>of</strong> 80 male Wistar rats (aged 8 weeks) were obtained<br />

from Charles River S.A. (Barcelona, Spain), with a mean<br />

weight <strong>of</strong> 249±23 g. Animals were kept in standard laboratory<br />

conditions (12/12 h light/darkness, 22 ±2°C room temperature,<br />

50–60% humidity) for at least 1 week before starting <strong>the</strong><br />

experiments. Animals were allowed access to tap water and rat<br />

chow ad libitum during this period. Animal experiments were<br />

licensed by Portuguese General Directorate <strong>of</strong> Veterinary Medicine<br />

(DGV). Housing and experimental treatment <strong>of</strong> animals<br />

were in accordance to <strong>the</strong> Guide for <strong>the</strong> Care and Use <strong>of</strong><br />

Laboratory Animals from <strong>the</strong> Institute for Laboratory Animal<br />

Research (ILAR 1996). The experiments complied with <strong>the</strong><br />

current laws <strong>of</strong> Portugal.<br />

Experimental protocol<br />

Each animal was individually housed during <strong>the</strong> experimental<br />

period in a polypropylene cage with a stainless-steel net<br />

at <strong>the</strong> top and wood chips at <strong>the</strong> screen bottom. Tap water and<br />

rat chow were given ad libitum during <strong>the</strong> entire experiment.<br />

Treatments in all groups were always conducted between 8:00<br />

and 10:00 AM. The administrations <strong>of</strong> vehicle (0.9% NaCl), PQ,<br />

and NaSAL were all made intraperitoneally (ip) in an injection<br />

volume <strong>of</strong> 0.5 mL/250 g <strong>of</strong> body weight. The schedule <strong>of</strong><br />

NaSAL administration (2 h after PQ) was chosen taking <strong>into</strong><br />

account <strong>the</strong> estimated average arrival time <strong>of</strong> <strong>the</strong> patient to <strong>the</strong><br />

hospital, after PQ <strong>into</strong>xication. The experimental dose <strong>of</strong><br />

NaSAL was chosen according to literature data <strong>of</strong> in vivo<br />

studies [28]. The PQ-administered dose is known to produce<br />

severe lung toxicity and death in rats within a few days [28–30].<br />

Each group was treated as described in Fig. 1. Briefly: (i)<br />

control group, n=8: animals were first administered with 0.9%<br />

NaCl. Animals were administered with one more administration<br />

<strong>of</strong> 0.9% NaCl 2 h later and sacrificed 24 h after <strong>the</strong> second<br />

injection. (ii) NaSAL group, n=24: animals were first<br />

administered with 0.9% NaCl. Animals were treated with one<br />

administration <strong>of</strong> NaSAL (200 mg/kg) 2 h later and sacrificed<br />

24 h (n=8, NaSAL 24 h group), 48 h (n=8, NaSAL 48 h group),<br />

and 96 h (n=8, NaSAL 96 h group) after <strong>the</strong> second injection.<br />

(iii) PQ group, n=24: animals were first <strong>into</strong>xicated with PQ<br />

(25 mg/kg). Animals were administered with one more<br />

administration <strong>of</strong> 0.9% NaCl 2 h later and sacrificed 24 h<br />

(n=8, PQ 24 h group), 48 h (n=8, PQ 48 h group), and 96 h<br />

(n=8, PQ 96 h group) after <strong>the</strong> second injection. (iv) PQ+<br />

NaSAL group, n=24: animals were first <strong>into</strong>xicated with PQ<br />

(25 mg/kg). Two hours later, animals were treated with NaSAL<br />

(200 mg/kg) and sacrificed 24 h (n=8, PQ+NaSAL 24 h<br />

group), 48 h (n=8, PQ+NaSAL 48 h group), and 96 h (n=8,<br />

PQ+NaSAL 96 h group) after <strong>the</strong> second injection.<br />

Surgical procedures<br />

Before sacrifice, anes<strong>the</strong>sia was <strong>induced</strong> with sodium thiopental<br />

(60 mg/kg, ip). In six rats <strong>of</strong> each group (nonhistological<br />

studies), lungs were perfused in situ through <strong>the</strong> pulmonary


artery with cold 0.9% NaCl for 3 min at a rate <strong>of</strong> 10 ml/min to<br />

remove most trapped blood volume. Simultaneous with <strong>the</strong><br />

perfusion initiation, left wall ventricle was cut to avoid cardiovascular<br />

volume overload. In <strong>the</strong> remaining two animals<br />

(histological study), <strong>the</strong> trachea was exposed and intubated.<br />

Lungs were inflated by administration <strong>of</strong> <strong>the</strong> fixative [4% (v/v)<br />

buffered formaldehyde; in situ fixation]. Cubic pieces were <strong>the</strong>n<br />

fixed by diffusion for 24 h and subsequently processed for<br />

routine paraffin histology. Serial sections (4 μm) <strong>of</strong> <strong>the</strong> paraffin<br />

blocks were cut by a microtome and mounted on slides coated<br />

with aminopropyl-triethoxysilane. The slides were dewaxed in<br />

xylene and hydrated through graded alcohols finishing in<br />

phosphate-buffered saline (10 mM PBS, pH 7.2).<br />

Tissue processing for nonhistological studies<br />

Lungs were removed, cleaned <strong>of</strong> all major cartilaginous<br />

tissues <strong>of</strong> <strong>the</strong> conducting airways, pat-dried with gauze, and<br />

processed as follows: right lungs (except <strong>the</strong> posterior and <strong>the</strong><br />

postcaval lobe) were homogenized (Ultra-Turrax homogenizer)<br />

in 2.5 ml <strong>of</strong> an ice-cold isotonic buffer (300 mM sucrose,<br />

10 mM Hepes, 2 mM EGTA, pH 7.4) followed by addition <strong>of</strong><br />

1 mM dithiothreitol (DTT), 1 mM phenylmethylsulfonyl<br />

fluoride (PMSF), and 5 μg/ml <strong>of</strong> each proteases inhibitor<br />

(pepstatin A, leupeptin, and aprotinin). Cytosolic fractions were<br />

prepared essentially as described by Atlante et al. [31] with<br />

slight modifications. Homogenates were centrifuged (600g,<br />

4°C, for 10 min) to remove <strong>the</strong> nuclei and unbroken cells. The<br />

resulting supernatants were <strong>the</strong>n centrifuged (9500g, 4°C, for<br />

10 min). Supernatants (cytosolic fraction) were recovered and<br />

stored (–80°C) until processed for Cyt c quantification. DNA<br />

was extracted from <strong>the</strong> posterior and postcaval lobe according to<br />

<strong>the</strong> standard method described by Shimelis et al. [32] with slight<br />

modifications. Briefly, lobes were homogenized (Ultra-Turrax<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

Fig. 1. Schematic representation <strong>of</strong> <strong>the</strong> administration protocols for <strong>the</strong> control, sodium salicylate (NaSAL), <strong>paraquat</strong> (PQ), and <strong>paraquat</strong> plus sodium salicylate (PQ+<br />

NaSAL) groups.<br />

homogenizer) in 2.5 ml <strong>of</strong> a lysis buffer (Tris-HCl 100 mM, pH<br />

8, EDTA 50 mM, 0.5% SDS). The homogenates were first<br />

incubated with RNase A (200 μg/ml, 2 h, 37°C) and<br />

subsequently with proteinase K (200 μg/ml, 2 h, 37°C). DNA<br />

was extracted twice with buffered phenol and with 1:1 mixture<br />

<strong>of</strong> buffered phenol:chlor<strong>of</strong>orm. For each extraction step, a mixture<br />

by inversion (5 min) followed by centrifugation (2000g, for<br />

10 min) to separate <strong>the</strong> phases was performed. The superior<br />

aqueous phase was recovered and DNA was precipitated by<br />

adding 0.1 vol <strong>of</strong> 3 M sodium acetate, pH 5.2, and 3 vol <strong>of</strong> icecold<br />

ethanol. The tubes were inverted for 5 min and <strong>the</strong> DNA<br />

was pelleted by centrifugation (13,000g, 4°C, for 30 min).<br />

Thereafter, DNA pellet was washed twice with 70% ethanol<br />

(4°C) to remove salt, air-dried overnight at 4°C (<strong>the</strong> Eppendorfs<br />

were left open but covered with aluminum foil during this<br />

procedure), and <strong>the</strong>n redissolved in 0.5 ml <strong>of</strong> TE buffer (Tris-<br />

HCl 10 mM, 1 mM EDTA, pH 7.6). Left lungs were used for<br />

preparation <strong>of</strong> cytoplasmic and nuclear extracts. Briefly, left<br />

lungs were homogenized (Ultra-Turrax homogenizer) in a AC<br />

buffer (cell lysis buffer:1 g <strong>of</strong> tissue/3 ml) containing 10 mM<br />

Hepes (pH 7.9), 10 mM KCl, 1.5 mM MgCl2, 0.2% igepal,<br />

0.5 mM EDTA, 0.1 mM EGTA, 1 mM DTT, and 0.25 mM<br />

PMSF and incubated on ice for 15 min. After a brief vortexing,<br />

<strong>the</strong> lysates were centrifuged (850g, 4°C for 10 min). The<br />

supernatants (cytoplasmic extracts) were saved and <strong>the</strong> pellets<br />

were resuspended (washing step) in 500 μl <strong>of</strong> AC buffer and<br />

incubated for 15 min on ice and <strong>the</strong>n centrifuged (14,000g, 4°C<br />

for 30 s). The supernatants (cytoplasmic extracts) were added to<br />

those obtained in <strong>the</strong> previous step, divided <strong>into</strong> aliquots, and<br />

stored at −80°C for posterior determination <strong>of</strong> caspases-1, -8,<br />

and -3 activities. The pellets were resuspended in 500 μl <strong>of</strong>BC<br />

buffer (nuclei lysis buffer) containing 20 mM Hepes, pH 7.9,<br />

420 mM NaCl, 1.5 mM MgCl2, 0.2% igepal, 0.5 mM EDTA,<br />

20% glycerol, 1 mM DTT, 0.25 mM PMSF, aprotinin (5 μg/ml),<br />

51


52 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

pepsatin (5 μg/ml), and leupeptin (5 μg/ml) and incubated on<br />

ice for 30 min. After a brief vortexing, <strong>the</strong> lysates were centrifuged<br />

(14,000g, 4°C, for 10 min). Supernatants (nuclear extracts)<br />

were collected, divided <strong>into</strong> aliquots, and stored at –80°C<br />

for semiquantification <strong>of</strong> p53 and AP-1 by fluorescent electrophoretic<br />

mobility shift assay (fEMSA).<br />

Oligonucleotides and DNA annealing<br />

Double-stranded fluorescent targets were prepared by mixing<br />

equimolar amounts <strong>of</strong> <strong>the</strong> two complementary singlestranded<br />

oligonucleotides (p53-FW-Cy5 or p53-FW with p53-<br />

R, AP-1-FW-Cy5 or AP-1-FW with AP-1-R, and NF-κB-FW<br />

with NF-κB-R) as described previously [28].<br />

Protein quantification<br />

Protein quantification was performed accordingly to <strong>the</strong><br />

method <strong>of</strong> Lowry et al. [33], using bovine serum albumin as<br />

standard.<br />

Semiquantification <strong>of</strong> transcriptional activation <strong>of</strong> lung nuclear<br />

proteins by fluorescent electrophoretic mobility shift assay<br />

The p53- and AP-1-binding assays and respective analysis<br />

were performed according to a previously reported method<br />

based on <strong>the</strong> binding <strong>of</strong> <strong>the</strong> transcription factors to <strong>the</strong>ir specific<br />

DNA recognition sequences [28]. Specificity <strong>of</strong> <strong>the</strong> DNA–<br />

protein complex was confirmed by <strong>the</strong> addition <strong>of</strong> a 50-fold<br />

excess <strong>of</strong> ei<strong>the</strong>r unlabeled specific competitor (SC, specific<br />

probe without <strong>the</strong> Cy5 label) or unlabeled nonspecific competitor<br />

(UC, which was <strong>the</strong> NF-κB unlabeled oligonucleotide,<br />

ei<strong>the</strong>r for AP-1 or p53).<br />

Quantification <strong>of</strong> caspase activities<br />

The enzymatic activities <strong>of</strong> caspases-1, -8, and -3 in lung<br />

tissues were evaluated using <strong>the</strong> commercially available caspase-1,<br />

-8, and -3 colorimetric substrates. Briefly, samples were<br />

adequately diluted (100 μg protein/well) in buffer (25 mM<br />

Hepes, pH 7.4, 0.1 Chaps, 10% sacarose, supplemented with<br />

10 mM DTT). Triplicate samples were incubated for 90 min in<br />

<strong>the</strong> dark at 37°C with 40 μM <strong>of</strong> each specific substrate. The<br />

cleavage <strong>of</strong> <strong>the</strong> substrate peptide by <strong>the</strong> respective caspases<br />

releases <strong>the</strong> chromophore p-nitroanilide, which is quantified<br />

spectrophotometrically at 405 nm. The enzymatic activities <strong>of</strong><br />

caspases-1, -8, and -3 in lung tissue homogenates were expressed<br />

as absorbance units (Uabs)/100 μg protein.<br />

Measurement <strong>of</strong> cytochrome c translocation<br />

The levels <strong>of</strong> Cyt c in <strong>the</strong> cytosolic fraction (50 μg protein)<br />

were analyzed by Western blot on 12% SDS-polyacrylamide gel<br />

under constant current (14–15 mA) according to <strong>the</strong> conventional<br />

methods partially modified by Fuentes et al. [34]. Briefly,<br />

separated proteins were electrotransferred (250 mA for 60 min)<br />

to PVDF (polyvinylidene difluoride) membranes using a Mini<br />

Trans-Blot Cell apparatus (Bio-Rad). The blots were blocked<br />

with 10% nonfat dried milk in TTBS (10 mM Tris/HCl, pH 7.5,<br />

150 mM NaCl, and 0.2% Tween 20) at 4°C overnight and <strong>the</strong>n<br />

incubated with primary antibody (diluted 1:1000 in TTBS+5%<br />

nonfat dried milk) for 1 to 2 h at room temperature. After<br />

washing (two times 5 min with TTBS), membranes were incubated<br />

(60 min at room temperature) with peroxidase-conjugated<br />

secondary antibodies (1:5000 in TTBS with 10% nonfat dried<br />

milk). After washing (2×5 min and 1×10 min), bound<br />

antibodies were visualized by chemiluminescence using <strong>the</strong><br />

ECL-Plus reagent.<br />

DNA fragmentation analysis by electrophoresis<br />

DNA concentrations were evaluated by ultraviolet spectrophotometry<br />

at 260 nm. Ten micrograms <strong>of</strong> DNA aliquots were<br />

electrophoresed on 2% agarose gel at 70±2 V for 2 h and<br />

stained with 0.8 μg/ml <strong>of</strong> ethidium bromide. For visualization <strong>of</strong><br />

apoptotic alterations, DNA bands were observed on a transilluminator<br />

and recorded in photographs.<br />

DNA fragmentation analysis by deoxynucleotidyl<br />

transferase-mediated deoxyuridinetriphosphate (dUTP) nick<br />

end-labeling assay<br />

DNA strand breaks <strong>of</strong> <strong>the</strong> lung tissue were analyzed qualitatively<br />

and semiquantitatively, by fluorescence and light<br />

microscopy, using a TUNEL assay commercial kit (In SituCell<br />

Death Detection Kit; Roche Molecular Biochemicals, Germany),<br />

with slight modifications introduced by Correia-da-Silva<br />

et al. [35]. Briefly, deparaffinized sections were pretreated with<br />

proteinase K (20 μg/ml) in 0.05 M Tris/HCl, pH 7.6 (30 min at<br />

37°C), to break up membranes and free DNA, and <strong>the</strong>n washed<br />

in PBS solution. The sections were dried and incubated (1 h, in a<br />

humidified chamber, at 37°C) with a reaction mixture containing<br />

terminal deoxynucleotidyl transferase and fluorescein-labeled<br />

deoxyuridinetriphosphate. Fluorescence photos were taken at<br />

this point. In light microscopy, incorporated fluorescein was<br />

detected by an anti-fluorescein-antibody conjugated with alkaline<br />

phosphatase. The slides were washed with PBS and<br />

incubated for 25 min in alkaline phosphatase substrate solution.<br />

The reaction was stopped with tap water and slides were<br />

counterstained with Mayer's hematoxylin solution (diluted 1:2),<br />

and mounted in aqueous medium (Aquatex; Merck, Darmstadt,<br />

Germany). TUNEL-positive cells were identified by <strong>the</strong><br />

presence <strong>of</strong> red reactivity. Negative controls were prepared<br />

without TdTenzyme and sections previously treated with DNase<br />

I (100 U; Roche Molecular Biochemicals) were used as positive<br />

controls. TUNEL-positive cells (TPC; green points at fluorescence<br />

microscopy) were counted/field (magnification 100×)<br />

using 5 slides/group and results recorded in a blinded fashion by<br />

an experienced histologist.<br />

Statistical analysis<br />

Results are expressed as mean ±SE (standard error). Statistical<br />

comparison between groups was estimated using <strong>the</strong>


nonparametric method <strong>of</strong> Kruskal-Wallis followed by Dunn's<br />

test. In all cases, P values lower than 0.05 were considered<br />

statistically significant.<br />

Results<br />

DNA laddering analysis<br />

Apoptosis can be measured by visualizing <strong>the</strong> fragmentation<br />

<strong>of</strong> nuclear DNA resulting in <strong>the</strong> appearance <strong>of</strong> incrementally<br />

sized low-molecular-weight DNA bands on ethidium bromidestained<br />

agarose gels (DNA ladder). In our work, electrophoretic<br />

analysis <strong>of</strong> DNA extracted from whole lung tissue <strong>of</strong> rats<br />

exhibited marked DNA fragmentation in <strong>the</strong> PQ 48 and 96 h<br />

groups (Fig. 2). An increase <strong>of</strong> smear in <strong>the</strong> lung samples <strong>of</strong> <strong>the</strong><br />

PQ 24 h group was evident, which might be indicative <strong>of</strong> an<br />

increase <strong>of</strong> DNA fragmentation. In contrast, in <strong>the</strong> control,<br />

NaSAL 24, 48, and 96 h, and in <strong>the</strong> PQ 24 h groups no DNA<br />

laddering was detected. The posttreatment with NaSAL<br />

completely prevented PQ-<strong>induced</strong> DNA fragmentation.<br />

TUNEL analysis<br />

The TUNEL assay detects nuclear DNA fragmentation by<br />

labeling free 3′-OH terminals with dUTP by TdT catalysis<br />

(Figs. 3A, 3B, and 3C). TUNEL analysis from whole lung tissue<br />

<strong>of</strong> control and only NaSAL-treated groups revealed very few<br />

TPC. Animals from PQ groups exhibited a time-dependent<br />

marked DNA fragmentation. The posttreatment with NaSAL <strong>of</strong><br />

<strong>the</strong> PQ-exposed animals significantly reduced PQ-<strong>induced</strong><br />

DNA fragmentation. The green fluorescent points in Fig. 3A<br />

correspond mainly to type I and II cell nuclei as suggested by<br />

Fig. 2. Representative electrophoretic analysis <strong>of</strong> <strong>the</strong> DNA ladder formation in<br />

whole lungs <strong>of</strong> rats from control, sodium salicylate (NaSAL), <strong>paraquat</strong> (PQ),<br />

and <strong>paraquat</strong> plus sodium salicylate (PQ+NaSAL) groups at three different<br />

sample times. Mr indicates <strong>the</strong> lane <strong>of</strong> <strong>the</strong> molecular weight marker. These<br />

experiments were repeated using six different lung homogenates with comparable<br />

results.<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

immunohistochemistry analysis (Fig. 3B). In Fig. 3C, <strong>the</strong><br />

semiquantification <strong>of</strong> <strong>the</strong> TPC obtained at <strong>the</strong> fluorescence<br />

TUNEL assay clearly shows <strong>the</strong> apoptotic effect <strong>of</strong> PQ and <strong>the</strong><br />

healing provided by NaSAL posttreatment.<br />

Enzymatic activities <strong>of</strong> <strong>the</strong> caspases-1, -8, and -3<br />

The enzymatic activities <strong>of</strong> <strong>the</strong> caspases-1, -8, and -3 in lungs<br />

are presented in Fig. 4. Animals from <strong>the</strong> PQ group exhibited a<br />

significant rise <strong>of</strong> <strong>the</strong> activities <strong>of</strong> both caspase-8 and caspase-3<br />

in lung tissue 24, 48, and 96 h post-PQ exposure, compared<br />

with animals from control and NaSAL groups. On <strong>the</strong> o<strong>the</strong>r<br />

hand, a statistically significant decrease was observed in <strong>the</strong><br />

activity <strong>of</strong> caspase-1 in <strong>the</strong> lungs <strong>of</strong> <strong>the</strong> PQ 24, 48, and 96 h<br />

groups, compared to control and NaSAL groups. The posttreatment<br />

with NaSAL <strong>of</strong> rats exposed to PQ completely<br />

reverted <strong>the</strong> PQ-mediated effects to all studied caspases toward<br />

values similar to those obtained for NaSAL groups. Also,<br />

noteworthy was <strong>the</strong> increase <strong>of</strong> caspase-1 in <strong>the</strong> NaSAL 24 and<br />

48 h groups, caspase-8 in <strong>the</strong> NaSAL 24, 48, and 96 h groups,<br />

and caspase-3 in <strong>the</strong> NaSAL 24 and 48 h groups compared to<br />

<strong>the</strong> respective control groups.<br />

Determination <strong>of</strong> cytochrome c concentrations<br />

Cyt c is an important apoptogenic factor in <strong>the</strong> intrinsic<br />

apoptotic pathway [13]. We observed a significant increase <strong>of</strong><br />

cytosolic Cyt c concentration as a consequence <strong>of</strong> PQ exposure,<br />

with <strong>the</strong> maximal induction observed 48–96 h later in<br />

comparison to control and NaSAL groups (Fig. 5), suggesting<br />

that <strong>the</strong> intrinsic pathway is involved in PQ apoptosis. Statistically<br />

significant decreases in cytosolic Cyt c concentrations<br />

were observed in animals <strong>of</strong> NaSAL 24, 48, and 96 h groups<br />

compared to control group. Of note was <strong>the</strong> significant decrease<br />

<strong>of</strong> cytosolic Cyt c <strong>of</strong> PQ+NaSAL groups in comparison not<br />

only to PQ-exposed but also to control groups.<br />

Activation <strong>of</strong> AP-1<br />

fEMSA was performed to study <strong>the</strong> effects <strong>of</strong> PQ in <strong>the</strong> rat<br />

lung expression <strong>of</strong> AP-1 and p53. As shown in Fig. 6A, PQ<br />

<strong>induced</strong> a significant time-dependent activation <strong>of</strong> AP-1 in rat<br />

lungs (Lanes 2–4) compared to control (Lane 1) and NaSAL<br />

groups (Lanes 8–10). The AP-1-binding activity appeared to<br />

result from <strong>the</strong> formation <strong>of</strong> a single complex or complexes <strong>of</strong><br />

very similar mobility. Only barely detectable expression levels<br />

<strong>of</strong> AP-1 were observed in whole lung nuclear extracts from <strong>the</strong><br />

control group. Noteworthy, was also <strong>the</strong> significant reduction <strong>of</strong><br />

AP-1 lung activation in <strong>the</strong> NaSAL 24, 48, and 96 h groups<br />

relative to control, <strong>the</strong> signal being completely absent in <strong>the</strong>se<br />

groups. Concerning <strong>the</strong> PQ+NaSAL 24, 48, and 96 groups<br />

(Lanes 11, 12, and 13, respectively), NaSAL treatment resulted<br />

in a significant reduction <strong>of</strong> PQ-<strong>induced</strong> AP-1 activation, <strong>the</strong><br />

AP-1 expression being similar to that <strong>of</strong> <strong>the</strong> control group. The<br />

specificity <strong>of</strong> <strong>the</strong> DNA–protein complex was confirmed in <strong>the</strong><br />

PQ 96 h group by <strong>the</strong> persistence <strong>of</strong> <strong>the</strong> bands in <strong>the</strong> competition<br />

experiment with a 50-fold molar excess <strong>of</strong> <strong>the</strong> UC (Lane 6) and<br />

53


54 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61


Fig. 4. Activity <strong>of</strong> <strong>the</strong> caspases-1, -8, and -3 in <strong>the</strong> control, sodium salicylate<br />

(NaSAL), <strong>paraquat</strong> (PQ), and <strong>paraquat</strong> plus sodium salicylate (PQ+NaSAL)<br />

groups at three different sampled times. Data are expressed as absorbance units<br />

(U abs)/100 μg protein. Values are given as mean±SE (n=6). a P


56 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

Fig. 5. Immunoblot analysis <strong>of</strong> <strong>the</strong> cytochrome c (Cyt c) release from rat lung mitochondria <strong>into</strong> <strong>the</strong> cytosol <strong>of</strong> <strong>the</strong> control (C), sodium salicylate (NaSAL), <strong>paraquat</strong><br />

(PQ), and <strong>paraquat</strong> plus sodium salicylate (PQ+NaSAL) groups. Blot is representative <strong>of</strong> six independent experiments. α-Tubulin Western blot is included as a loading<br />

protein control. Values are given as mean±SE (n=6). a P


caspase-1 activity observed in rats <strong>of</strong> PQ 24, 48, and 96 h<br />

groups could be <strong>the</strong> consequence <strong>of</strong> antiapoptotic IL-1β effects.<br />

These data are <strong>of</strong> particular relevance to neutrophils, because<br />

<strong>the</strong>y express caspase-1 [42] and exhibit delayed apoptosis<br />

following exposure to exogenous IL-1β [43] or to LPS, <strong>the</strong><br />

latter effect being in part mediated via autocrine production <strong>of</strong><br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

Fig. 6. Representative fEMSA gel views <strong>of</strong> activator protein-1 (AP-1) (A) and p53 (B) activation <strong>induced</strong> by PQ in lungs at 24, 48, and 96 h. Nuclear extracts from <strong>the</strong><br />

different groups were prepared and subjected to fEMSA as described under Materials and Methods. Lane 1, control group; Lane 2, PQ 24 h; Lane 3, PQ 48 h; Lane 4,<br />

PQ 96 h; Lane 5, blank; Lane 6, competition experiment with a 50-fold molar excess <strong>of</strong> a nonspecific competitor (UC) compared to specific probe (SP); Lane 7,<br />

competition experiment with a 50-fold molar excess <strong>of</strong> a specific competitor (SC, unlabeled specific probe) compared to SP; Lanes 8, 9, and 10, sodium salicylate<br />

(NaSAL) 24, 48, and 96 h groups, respectively; Lanes 11, 12, and 13, PQ+NaSAL 24, 48, and 96 h groups, respectively. The positions <strong>of</strong> specific AP-1/DNA-binding<br />

complexes (bands 1–3) and p53/DNA-binding complexes are indicated in A and B, respectively. NS band represents nonspecific binding. The localization <strong>of</strong> <strong>the</strong> free<br />

probe (FP) is also indicated. The results presented in A and B are representative <strong>of</strong> six independent experiments.<br />

IL-1β [42]. This process is also important for <strong>the</strong> normal resolution<br />

<strong>of</strong> inflammation in tissues, because it leads to recognition<br />

and clearance <strong>of</strong> <strong>the</strong> apoptotic neutrophils by macrophages [44].<br />

Inhibition <strong>of</strong> caspase-1 activity as <strong>the</strong> consequence <strong>of</strong> PQ<br />

exposure might thus imply higher recruitment and permanence<br />

<strong>of</strong> inflammatory cells, namely neutrophils, which will not be<br />

57


58 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

removed from lungs by apoptosis. One <strong>of</strong> <strong>the</strong> main and<br />

interesting results concerning <strong>the</strong> inclusion <strong>of</strong> NaSAL in <strong>the</strong><br />

<strong>the</strong>rapy <strong>of</strong> rats <strong>into</strong>xicated by PQ was <strong>the</strong> increase <strong>of</strong> caspase-1<br />

activity compared to PQ-only exposed and control groups. We<br />

have previously shown, both by assessing <strong>the</strong> MPO activity and<br />

by histopathological studies, <strong>the</strong> widespread infiltration <strong>of</strong><br />

neutrophils and macrophages in <strong>the</strong> lungs <strong>of</strong> PQ-only exposed<br />

rats [28–30] as well as its reduction in PQ+NaSAL groups [28].<br />

Considering <strong>the</strong> involvement <strong>of</strong> caspase-1 in neutrophil<br />

apoptosis, our results suggest that <strong>the</strong> increase <strong>of</strong> caspase-1<br />

activity is a possible explanation for <strong>the</strong> decrease <strong>of</strong> neutrophil<br />

lung infiltration in PQ+NaSAL groups [28].<br />

In most cells, <strong>the</strong> morphological and biochemical features <strong>of</strong><br />

apoptosis seem to be associated with <strong>the</strong> cleavage <strong>of</strong> genomic<br />

DNA <strong>into</strong> large fragments and later <strong>into</strong> oligonucleosomal<br />

fragments by a nuclease which is activated exclusively in<br />

apoptosis [45]. With reference to our electrophoretic results<br />

(Fig. 2), we only observed DNA fragmentation in PQ 48 and<br />

96 h groups (a characteristic ladder-like pattern <strong>of</strong> DNA),<br />

although an increase <strong>of</strong> smear was already observed in <strong>the</strong> PQ<br />

24 group. Using TUNEL assay techniques, it was also possible<br />

to observe a marked increase <strong>of</strong> apoptosis <strong>of</strong> lung cells as a<br />

consequence <strong>of</strong> PQ exposure, in a time-dependent manner<br />

(Figs. 3A, 3B, and 3C), as well as its remission by NaSAL. The<br />

reason why <strong>the</strong> activation <strong>of</strong> caspase-8 (24, 48, and 96 h),<br />

caspase-3 (24 and 48 h), and caspase-1 (24 and 48 h) in only<br />

NaSAL-treated groups did not induce apoptosis <strong>of</strong> lung cells<br />

visualized by DNA laddering and/or TUNEL assay is not clear.<br />

Although it is a widely accepted concept that activation <strong>of</strong><br />

caspase-3 marks <strong>the</strong> “point <strong>of</strong> no return” in <strong>the</strong> pathway to<br />

apoptotic death <strong>of</strong> mammalian cells, our results could not<br />

corroborate that. Noteworthy, Guthmann et al. [39] did not find<br />

apoptosis, ei<strong>the</strong>r in lung tissue or in freshly isolated type II cells<br />

in response to sublethal hyperoxia despite <strong>the</strong> significant activation<br />

<strong>of</strong> caspases-8 and -3. They concluded that <strong>the</strong> increase <strong>of</strong><br />

caspases-8 and -3, in response to sublethal hyperoxia, did not<br />

mark <strong>the</strong> point <strong>of</strong> no return. The same conclusion could be<br />

ascertained considering our results. This interpretation <strong>of</strong> our<br />

findings is also strongly corroborated by Perfettini and Kroemer<br />

[46], who postulated that caspase activation is not synonymous<br />

<strong>of</strong> apoptotic demise. In addition, it could not be excluded that<br />

<strong>the</strong> extent <strong>of</strong> caspase-3 activation, 1.23- and 1.29-fold in <strong>the</strong><br />

NaSAL 24 and 48 h groups, respectively, in comparison with<br />

control group, is not high enough to induce apoptosis, although<br />

a 1.62-, 1.78-, and 1.96-fold increase <strong>of</strong> caspase-3 in <strong>the</strong> PQ 24,<br />

48, and 96 h groups, respectively, in comparison with control<br />

group, results in apoptosis in lung cells. Thus, we hypo<strong>the</strong>size<br />

that caspase-3 activation is <strong>the</strong>n followed by apoptosis, when its<br />

activation occurs via strong mitochondrial damage resulting in<br />

Cyt c release in PQ-<strong>induced</strong> pulmonary ROS toxicity. This<br />

concept is supported by recent findings showing that mitochondrial<br />

Cyt c release is a key event in hyperoxia-<strong>induced</strong> lung<br />

injury [47].<br />

By using fEMSA, we demonstrated that PQ exposure led to<br />

an increase <strong>of</strong> AP-1 DNA binding in lungs (Fig. 6A). Similar<br />

results were also demonstrated in vitro by Chen and Sun [48]<br />

and Li and Sun [49] using PC12 cells, and by Zhou et al. [50] in<br />

skeletal muscle cells. These results are, to some extent, similar<br />

to <strong>the</strong> effects observed for NF-κB [28], although with slower<br />

kinetics, which is consistent with <strong>the</strong> mode <strong>of</strong> activation<br />

requiring de novo syn<strong>the</strong>sis <strong>of</strong> fos and jun subunits [25]. In<br />

accordance with data reported in <strong>the</strong> literature, where NaSAL<br />

proved to inhibit <strong>the</strong> transactivation <strong>of</strong> AP-1 [51], we also<br />

observed that NaSAL itself reduces AP-1 activation. In<br />

agreement, NaSAL, given 2 h after PQ, inhibited <strong>the</strong> increase<br />

<strong>of</strong> AP-1 DNA-binding activity <strong>induced</strong> by PQ. Since <strong>the</strong><br />

promoter regions <strong>of</strong> many inflammatory cytokines and chemokines<br />

(e.g., TNF-α and IL-1β) contain AP-1-binding sites<br />

[52], <strong>the</strong> inhibition <strong>of</strong> AP-1 activation by NaSAL contributes to<br />

protect lungs against oxidative stress-<strong>induced</strong> inflammation.<br />

Moreover, AP-1 is involved in <strong>the</strong> regulation <strong>of</strong> antioxidant<br />

enzymes by <strong>the</strong> presence <strong>of</strong> AP-1-response elements in <strong>the</strong><br />

promoter regions <strong>of</strong> genes encoding glutathione peroxidase<br />

(GPx) and catalase (CAT) [53]. We previously observed that<br />

NaSAL attenuates <strong>the</strong> increase <strong>of</strong> CAT and GPx activities near<br />

to <strong>the</strong> levels <strong>of</strong> control and NaSAL groups and that it was<br />

possible to establish a relationship between NF-κB expression<br />

and CAT and GPx activities [28]. The same correlation could be<br />

ascertained for AP-1. Never<strong>the</strong>less, <strong>the</strong> question <strong>of</strong> whe<strong>the</strong>r or<br />

not AP-1 activation plays an essential role on PQ-<strong>induced</strong><br />

apoptosis still needs to be addressed.<br />

Besides AP-1, PQ also <strong>induced</strong> an increase <strong>of</strong> p53<br />

expression in lungs (Fig. 6B). In accordance with our results,<br />

it was previously demonstrated that ROS play several distinct<br />

roles in <strong>the</strong> p53 pathway, such as being important activators <strong>of</strong><br />

p53 expression through <strong>the</strong>ir capacity to induce DNA strand<br />

breaks, and also by regulating <strong>the</strong> DNA binding <strong>of</strong> p53 [26],<br />

since p53 protein contains a DNA-binding domain structure that<br />

depends on <strong>the</strong> binding <strong>of</strong> zinc to critical redox-sensitive<br />

cysteines [54]. In agreement with AP-1 discussion, GPx activity<br />

is also transcriptionally activated by p53 [55]. Our findings are<br />

corroborated by <strong>the</strong> results <strong>of</strong> previous studies designed to<br />

investigate <strong>the</strong> role <strong>of</strong> p53 in <strong>the</strong> progression <strong>of</strong> PQ-<strong>induced</strong><br />

apoptosis [56]. These authors used two cell lines, wild-type<br />

p53-expressing human lung epi<strong>the</strong>lial-like cell line (L132) and a<br />

p53-deficient human promyelocytic leukemia cell line (U937),<br />

and explored <strong>the</strong> linkage among p53, DNA damage, and<br />

apoptosis. Following PQ exposure <strong>of</strong> L132 cells, <strong>the</strong> percentage<br />

<strong>of</strong> S-phase cells decreased significantly and <strong>the</strong> expression <strong>of</strong><br />

p53 protein increased, suggesting that entry <strong>into</strong> S-phase from<br />

G1-phase was blocked. U937 cells showed complete resistance<br />

to PQ. Those results suggested that PQ-<strong>induced</strong> DNA damage<br />

caused G1 arrest and apoptosis only in L132 cells, and that p53<br />

protein accumulation was required for <strong>the</strong> induction <strong>of</strong><br />

apoptosis by PQ. In addition, TNF-α, which is <strong>induced</strong> by<br />

PQ (see above) has also been described as enhancing p53<br />

mRNA expression, through <strong>the</strong> induction <strong>of</strong> <strong>the</strong> NF-κB [57].<br />

On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> inclusion <strong>of</strong> NaSAL in <strong>the</strong> <strong>the</strong>rapeutic<br />

regime <strong>of</strong> PQ-<strong>into</strong>xicated rats decreased PQ-<strong>induced</strong> p53<br />

expression. This might be <strong>the</strong> result <strong>of</strong> salicylates being<br />

important hydroxyl radical (HO . ) scavengers [58]. Our hypo<strong>the</strong>sis<br />

is that HO . works as a messenger for <strong>the</strong> activation <strong>of</strong> this<br />

tumor suppressor protein. Never<strong>the</strong>less, <strong>the</strong> slight increase <strong>of</strong><br />

p53 expression observed in NaSAL 24 and 48 h groups


compared to control should not be dismissed. It might be <strong>the</strong><br />

result <strong>of</strong> <strong>the</strong> NF-κB inhibition [28], since NF-κB inhibition has<br />

been correlated to <strong>the</strong> enhancement <strong>of</strong> p53 expression [59,60].<br />

In conclusion, our results demonstrate that PQ causes<br />

apoptosis by Cyt c release, increase <strong>of</strong> caspase-3 and -8 activity,<br />

decrease <strong>of</strong> caspase-1 activity, and increase <strong>of</strong> p53 and AP-1<br />

expression, resulting in DNA fragmentation (Fig. 7). Treatment<br />

with NaSAL <strong>of</strong> PQ-<strong>into</strong>xicated rats blocked to some extent<br />

<strong>the</strong>se events, with <strong>the</strong> consequent abolition <strong>of</strong> DNA fragmentation.<br />

In view <strong>of</strong> our results, it is plausible to conclude that a<br />

high-dose <strong>the</strong>rapy with NaSAL, starting as soon as possible<br />

after PQ <strong>into</strong>xication, may constitute a promising treatment <strong>of</strong><br />

PQ poisonings, not only as <strong>the</strong> consequence <strong>of</strong> <strong>the</strong> effective<br />

inhibition <strong>of</strong> proinflammatory factors such as NF-κB, scavenging<br />

ROS, inhibition <strong>of</strong> MPO, and inhibition <strong>of</strong> platelet<br />

aggregation [28], but also by its potential beneficial effects at<br />

<strong>the</strong> apoptotic pathways (Fig. 7). Despite <strong>the</strong> relevance <strong>of</strong> each<br />

beneficial effect, it seems logical that this results from a<br />

multiprotective action <strong>of</strong> NaSAL. This is important, since<br />

NaSAL was previously shown to protect lungs <strong>of</strong> PQchallenged<br />

rats, as confirmed by an amelioration <strong>of</strong> practically<br />

all toxicological parameters, which ended up in <strong>the</strong> achievement<br />

<strong>of</strong> full survival <strong>of</strong> <strong>the</strong> tested animals [28]. The present data<br />

reinforce <strong>the</strong> potential use <strong>of</strong> this interesting molecule in <strong>the</strong><br />

protection against PQ-<strong>induced</strong> lung damage.<br />

Acknowledgments<br />

Ricardo Dinis-Oliveira acknowledges FCT for his Ph.D.<br />

grant (SFRH/BD/13707/2003). The authors are thankful to<br />

Pr<strong>of</strong>essor Natércia from <strong>the</strong> Biochemistry Department <strong>of</strong> <strong>the</strong><br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

Fig. 7. Schematic illustration overview <strong>of</strong> <strong>the</strong> <strong>multiple</strong> toxic acute signals <strong>induced</strong> by <strong>paraquat</strong> in <strong>the</strong> lungs and <strong>the</strong> prevention obtained by sodium salicylate treatment.<br />

AP-1, activator protein-1; NF-κB, nuclear factor kappa-B; p53, tumor suppressor protein.<br />

Faculty <strong>of</strong> Pharmacy, University <strong>of</strong> Porto, for her precious help<br />

in <strong>the</strong> TUNEL experiments.<br />

References<br />

[1] Dinis-Oliveira, R. J.; Sarmento, A.; Reis, P.; Amaro, A.; Remião, F.;<br />

Bastos, M. L.; Carvalho, F. Acute <strong>paraquat</strong> poisoning: report <strong>of</strong> a survival<br />

case following intake <strong>of</strong> a potential lethal dose. Pediatr. Emerg. Care<br />

22:537–540; 2006.<br />

[2] Dinis-Oliveira, R. J.; Valle, M. J.; Bastos, M. L.; Carvalho, F.; Sanchez-<br />

Navarro, A. Kinetics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> isolated rat lung: influence <strong>of</strong><br />

sodium depletion. Xenobiotica 36:724–737; 2006.<br />

[3] Fabisiak, J. P.; Kagan, W. E.; Ritov, V. B.; Johnson, D. E.; Lazo, J. S. Bcl-2<br />

inhibits selective oxidation and externalization <strong>of</strong> phosphatidylserine<br />

during <strong>paraquat</strong>-<strong>induced</strong> apoptosis. Am. J. Physiol. Cell Physiol. 272:<br />

C675–C684; 1997.<br />

[4] Melchiorri, D.; Duca, C. D.; Piccirilli, S.; Trombetta, G.; Bagetta, G.;<br />

Nistico, G. Intrahippocampal injection <strong>of</strong> <strong>paraquat</strong> produces apoptotic cell<br />

death which is prevented by <strong>the</strong> lazaroid U74389G, in rats. Life Sci.<br />

62:1927–1932; 1998.<br />

[5] McCarthy, S.; Somayajulu, M.; Sikorska, M.; Borowy-Borowski, H.;<br />

Pandey, S. Paraquat induces oxidative stress and neuronal cell death;<br />

neuroprotection by water-soluble coenzyme Q10. Toxicol. Appl. Pharmacol.<br />

201:21–31; 2004.<br />

[6] Cappelletti, G.; Maggioni, M. G.; Maci, R. Apoptosis in human lung<br />

epi<strong>the</strong>lial cells: triggering by <strong>paraquat</strong> and modulation by antioxidants.<br />

Cell Biol. Int. 22:671–678; 1998.<br />

[7] Guinee, D. J.; Brambilla, E.; Fleming, M.; Hayashi, T.; Rahn, M.; Koss, M.;<br />

Ferrans, V.; Travis, W. The potential role <strong>of</strong> BAX and BCL-2 expression in<br />

diffuse alveolar damage. Am. J. Pathol. 151:999–1007; 1997.<br />

[8] Uhal, B. D. Cell cycle kinetics in <strong>the</strong> alveolar epi<strong>the</strong>lium. Am. J. Physiol.<br />

272:L1031–L1045; 1997.<br />

[9] Bismuth, C.; Hall, A. H., eds. Paraquat poisoning: <strong>mechanisms</strong>, prevention,<br />

treatment, vol. 10; 1995.<br />

[10] Raff, M. C. Social controls on cell survival and cell death. Nature<br />

356:397–400; 1992.<br />

59


60 R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

[11] Chin, Y. E.; Kitagawa, M.; Kuida, K.; Flavell, R. A.; Fu, X. Y. Activation<br />

<strong>of</strong> <strong>the</strong> STAT signaling pathway can cause expression <strong>of</strong> caspase 1 and<br />

apoptosis. Mol. Cell. Biol. 17:5328–5337; 1997.<br />

[12] Cohen, G. M. Caspases: <strong>the</strong> executioners <strong>of</strong> apoptosis. Biochem. J.<br />

326:1–16; 1997.<br />

[13] Hengartner, M. O. The biochemistry <strong>of</strong> apoptosis. Nature 407:770–776;<br />

2000.<br />

[14] Nunez, G.; Benedict, M. A.; Hu, Y.; Inohara, N. Caspases: <strong>the</strong> proteases <strong>of</strong><br />

<strong>the</strong> apoptotic pathway. Oncogene 17:3237–3245; 1998.<br />

[15] Nagata, S. Apoptosis by death factor. Cell 88:355–365; 1997.<br />

[16] Ashkenazi, A.; Dixit, V. M. Death receptors: signaling and modulation.<br />

Science 281:1305–1308; 1998.<br />

[17] Skulachev, V. P. Cytochrome c in <strong>the</strong> apoptotic and antioxidant cascades.<br />

FEBS Lett. 423:275–280; 1998.<br />

[18] Reed, J. C. Cytochrome c: can't live with it–can't live without it. Cell<br />

91:559–562; 1997.<br />

[19] Friedlander, R. M.; Gagliardini, V.; Rotello, R. J.; Yuan, J. Functional role<br />

<strong>of</strong> interleukin 1 beta (IL-1 beta) in IL-1 beta-converting enzyme-mediated<br />

apoptosis. J. Exp. Med. 184:717–724; 1996.<br />

[20] Rowe, S. J.; Allen, L.; Ridger, V. C.; Hellewell, P. G.; Whyte, M. K.<br />

Caspase-1-deficient mice have delayed neutrophil apoptosis and a<br />

prolonged inflammatory response to lipopolysaccharide-<strong>induced</strong> acute<br />

lung injury. J. Immunol. 169:6401–6407; 2002.<br />

[21] Abate, C.; Curran, T. Encounters with Fos and Jun on <strong>the</strong> road to AP-1.<br />

Semin. Cancer Biol. 1:19–26; 1990.<br />

[22] Smeyne, R. J.; Vendrell, M.; Hayward, M.; Baker, S. J.; Miao, G. G.;<br />

Schilling, K.; Robertson, L. M.; Curran, T.; M<strong>organ</strong>, J. I. Continuous c-fos<br />

expression precedes programmed cell death in vivo. Nature 363:166–169;<br />

1993.<br />

[23] Wisdom, R. AP-1: one switch for many signals. Exp. Cell Res.<br />

253:180–185; 1999.<br />

[24] Gomez del Arco, P.; Martinez-Martinez, S.; Calvo, V.; Armesilla, A. L.;<br />

Redondo, J. M. Antioxidants and AP-1 activation: a brief overview.<br />

Immunobiology 198:273–278; 1997.<br />

[25] Angel, P.; Karin, M. The role <strong>of</strong> Jun, Fos and <strong>the</strong> AP-1 complex in cellproliferation<br />

and transformation. Biochim. Biophys. Acta 1072:129–157;<br />

1991.<br />

[26] Meplan, C.; Richard, M. J.; Hainaut, P. Redox signalling and transition<br />

metals in <strong>the</strong> control <strong>of</strong> <strong>the</strong> p53 pathway. Biochem. Pharmacol. 59:25–33;<br />

2000.<br />

[27] Oren, M.; Damalas, A.; Gottlieb, T.; Michael, D.; Taplick, J.; Leal, J. F.;<br />

Maya, R.; Moas, M.; Seger, R.; Taya, Y.; Ben-Ze'Ev, A. Regulation <strong>of</strong> p53:<br />

intricate loops and delicate balances. Ann. N. Y. Acad. Sci. 973:374–383;<br />

2002.<br />

[28] Dinis-Oliveira, R. J.; Sousa, C.; Remião, F.; Duarte, J. A.; Sanchez-<br />

Navarro, A.; Bastos, M. L.; Carvalho, F. Full survival <strong>of</strong> <strong>paraquat</strong>-exposed<br />

rats after treatment with sodium salicylate. Free Radic. Biol. Med.<br />

42:1017–1028; 2007.<br />

[29] Dinis-Oliveira, R. J.; Duarte, J. A.; Remiao, F.; Sanchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. Single high dose dexamethasone treatment<br />

decreases <strong>the</strong> pathological effects and increases <strong>the</strong> survival rat <strong>of</strong><br />

<strong>paraquat</strong>-<strong>into</strong>xicated rats. Toxicology 227:73–85; 2006.<br />

[30] Dinis-Oliveira, R. J.; Remião, F.; Duarte, J. A.; Sanchez-Navarro, A.;<br />

Bastos, M. L.; Carvalho, F. P-glycoprotein induction: an antidotal pathway<br />

for <strong>paraquat</strong>-<strong>induced</strong> lung toxicity. Free Radic. Biol. Med. 41:1213–1224;<br />

2006.<br />

[31] Atlante, A.; Calissano, P.; Bobba, A.; Azzariti, A.; Marra, E.; Passarella, S.<br />

Cytochrome c is released from mitochondria in a reactive oxygen species<br />

(ROS)-dependent fashion and can operate as a ROS scavenger and as a<br />

respiratory substrate in cerebellar neurons undergoing excitotoxic death.<br />

J. Biol. Chem. 275:37159–37166; 2000.<br />

[32] Shimelis, O.; Zhou, X.; Li, G.; Giese, R. W. Phenolic extraction <strong>of</strong> DNA<br />

from mammalian tissues and conversion to deoxyribonucleoside-5′monophosphates<br />

devoid <strong>of</strong> ribonucleotides. J. Chromatogr. A<br />

1053:143–149; 2004.<br />

[33] Lowry, O. H. N.; Rosebrough, N. J.; Farr, A. L.; Randall, R. J. Protein<br />

measurement with Folin phenol reagent. J. Biol. Chem. 193:265–275;<br />

1951.<br />

[34] Fuentes, J. M.; Lompre, A. M.; Moller, J. V.; Falson, P.; le Maire, M. Clean<br />

Western blots <strong>of</strong> membrane proteins after yeast heterologous expression<br />

following a shortened version <strong>of</strong> <strong>the</strong> method <strong>of</strong> Perini et al. Anal. Biochem.<br />

285:276–278; 2000.<br />

[35] Correia-da-Silva, G.; Bell, S. C.; Pringle, J. H.; Teixeira, N. A. Patterns <strong>of</strong><br />

uterine cellular proliferation and apoptosis in <strong>the</strong> implantation site <strong>of</strong> <strong>the</strong> rat<br />

during pregnancy. Placenta 25:538–547; 2004.<br />

[36] Ishida, Y.; Takayasu, T.; Kimura, A.; Hayashi, T.; Kakimoto, N.;<br />

Miyashita, T.; Kondo, T. Gene expression <strong>of</strong> cytokines and growth factors<br />

in <strong>the</strong> lungs after <strong>paraquat</strong> administration in mice. Leg. Med. (Tokyo)<br />

8:102–109; 2006.<br />

[37] Kim, K. M.; Song, J. J.; An, J. Y.; Kwon, Y. T.; Lee, Y. J. Pretreatment <strong>of</strong><br />

acetylsalicylic acid promotes tumor necrosis factor-related apoptosisinducing<br />

ligand-<strong>induced</strong> apoptosis by down-regulating BCL-2 gene<br />

expression. J. Biol. Chem. 280:41047–41056; 2005.<br />

[38] Ramesh, G.; Reeves, W. B. Salicylate reduces cisplatin nephrotoxicity by<br />

inhibition <strong>of</strong> tumor necrosis factor-alpha. Kidney Int. 65:490–499; 2004.<br />

[39] Guthmann, F.; Wissel, H.; Schachtrup, C.; Tolle, A.; Rudiger, M.; Spener,<br />

F.; Rustow, B. Inhibition <strong>of</strong> TNFalpha in vivo prevents hyperoxiamediated<br />

activation <strong>of</strong> caspase 3 in type II cells. Respir. Res. 6:10; 2005.<br />

[40] Klampfer, L.; Cammenga, J.; Wisniewski, H. G.; Nimer, S. D. Sodium<br />

salicylate activates caspases and induces apoptosis <strong>of</strong> myeloid leukemia<br />

cell lines. Blood 93:2386–2394; 1999.<br />

[41] Tomita, M.; Okuyama, T.; Hidaka, K. Changes in mRNAs <strong>of</strong> inducible<br />

nitric oxide synthase and interleukin-1 beta in <strong>the</strong> liver, kidney and lung<br />

tissues <strong>of</strong> rats acutely exposed to <strong>paraquat</strong>. Leg. Med. (Tokyo) 1:127–134;<br />

1999.<br />

[42] Watson, R. W.; Rotstein, O. D.; Parodo, J.; Bitar, R.; Marshall, J. C. The<br />

IL-1 beta-converting enzyme (caspase-1) inhibits apoptosis <strong>of</strong> inflammatory<br />

neutrophils through activation <strong>of</strong> IL-1 beta. J. Immunol.<br />

161:957–962; 1998.<br />

[43] Colotta, F.; Re, F.; Polentarutti, N.; Sozzani, S.; Mantovani, A. Modulation<br />

<strong>of</strong> granulocyte survival and programmed cell death by cytokines and<br />

bacterial products. Blood 80:2012–2020; 1992.<br />

[44] Haslett, C. Granulocyte apoptosis and inflammatory disease. Br. Med.<br />

Bull. 53:669–683; 1997.<br />

[45] Oberhammer, F.; Wilson, J. W.; Dive, C.; Morris, I. D.; Hickman, J. A.;<br />

Wakeling, A. E.; Walker, P. R.; Sikorska, M. Apoptotic death in epi<strong>the</strong>lial<br />

cells: cleavage <strong>of</strong> DNA to 300 and/or 50 kb fragments prior to or in <strong>the</strong><br />

absence <strong>of</strong> internucleosomal fragmentation. EMBO J. 12:3679–3684;<br />

1993.<br />

[46] Perfettini, J. L.; Kroemer, G. Caspase activation is not death. Nat.<br />

Immunol. 4:308–310; 2003.<br />

[47] Pagano, A.; Donati, Y.; Metrailler, I.; Barazzone-Argir<strong>of</strong>fo, C. Mitochondrial<br />

cytochrome c release is a key event in hyperoxia-<strong>induced</strong> lung injury:<br />

protection by cyclosporin A. Am. J. Physiol. 286:L275–L283; 2004.<br />

[48] Chen, Y.; Sun, A. Y. Activation <strong>of</strong> transcription factor AP-1 by<br />

extracellular ATP in PC12 cells. Neurochem. Res. 23:543–550; 1998.<br />

[49] Li, X.; Sun, A. Y. Paraquat <strong>induced</strong> activation <strong>of</strong> transcription factor AP-1<br />

and apoptosis in PC12 cells. J. Neural Transm. 106:1–21; 1999.<br />

[50] Zhou, L. Z.; Johnson, A. P.; Rando, T. A. NF kappa B and AP-1 mediate<br />

transcriptional responses to oxidative stress in skeletal muscle cells. Free<br />

Radic. Biol. Med. 31:1405–1416; 2001.<br />

[51] Tegeder, I.; Niederberger, E.; Israr, E.; Guhring, H.; Brune, K.;<br />

Euchenh<strong>of</strong>er, C.; Grosch, S.; Geisslinger, G. Inhibition <strong>of</strong> NF-kappaB<br />

and AP-1 activation by R- and S-flurbipr<strong>of</strong>en. FASEB J. 15:595–597;<br />

2001.<br />

[52] Roebuck, K. A.; Carpenter, L. R.; Lakshminarayanan, V.; Page, S. M.;<br />

Moy, J. N.; Thomas, L. L. Stimulus-specific regulation <strong>of</strong> chemokine<br />

expression involves differential activation <strong>of</strong> <strong>the</strong> redox-responsive<br />

transcription factors AP-1 and NF-kappaB. J. Leukoc. Biol. 65:291–298;<br />

1999.<br />

[53] Jornot, L.; Junod, A. F. Hyperoxia, unlike phorbol ester, induces<br />

glutathione peroxidase through a protein kinase C-independent mechanism.<br />

Biochem. J. 326:117–123; 1997.<br />

[54] Hainaut, P.; Milner, J. Redox modulation <strong>of</strong> p53 conformation and<br />

sequence-specific DNA binding in vitro. Cancer Res. 53:4469–4473;<br />

1993.


[55] Tan, M.; Swaroop, S. L. M.; Guan, K.; Oberley, L. W.; Sun, Y.<br />

Transcriptional activation <strong>of</strong> <strong>the</strong> human glutathione peroxidase promoter<br />

by p53. J. Biol. Chem. 274:12061–12066; 1999.<br />

[56] Takeyama, N.; Tanaka, T.; Yabuki, T.; Nakatani, T. The involvement <strong>of</strong><br />

p53 in <strong>paraquat</strong>-<strong>induced</strong> apoptosis in human lung epi<strong>the</strong>lial-like cells. Int.<br />

J. Toxicol. 23:33–40; 2004.<br />

[57] Reisman, D.; Loging, W. T. Transcriptional regulation <strong>of</strong> <strong>the</strong> p53 tumor<br />

suppressor gene. Semin. Cancer Biol. 8:317–324; 1998.<br />

R.J. Dinis-Oliveira et al. / Free Radical Biology & Medicine 43 (2007) 48–61<br />

[58] van Jaarsveld, H.; Kuyl, J. M.; van Zyl, G. F.; Barnard, H. C. Salicylate in<br />

<strong>the</strong> perfusate during ischemia/reperfusion prevented mitochondrial injury.<br />

Res. Commun. Mol. Pathol. Pharmacol. 86:287–295; 1994.<br />

[59] Mayo, L. D.; Donner, D. B. The PTEN, Mdm2, p53 tumor suppressoroncoprotein<br />

network. Trends Biochem. Sci. 27:462–467; 2002.<br />

[60] Zhang, Z.; Ma, J.; Li, N.; Sun, N.; Wang, C. Expression <strong>of</strong> nuclear factorkappaB<br />

and its clinical significance in nonsmall-cell lung cancer. Ann.<br />

Thorac. Surg.243–248; 2006.<br />

61


________________________________________Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies<br />

1. INTEGRATED OVERVIEW OF THE PERFORMED STUDIES<br />

PART III<br />

1. INTEGRATED OVERVIEW OF THE PERFORMED STUDIES<br />

207


Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies________________________________________<br />

208


________________________________________Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies<br />

1. INTEGRATED OVERVIEW OF THE PERFORMED STUDIES<br />

Few years after introduction <strong>of</strong> PQ <strong>into</strong> <strong>the</strong> market it became clear that it was a<br />

serious hazard to humans as a consequence <strong>of</strong> its misuse. In early reports (Bullivant,<br />

1966; Campbell, 1968), accidental poisoning from drinking <strong>the</strong> dark brown concentrate,<br />

which resembled a cola drink after it has been decanted <strong>into</strong> s<strong>of</strong>t-drink bottles, was<br />

common. Nowadays, suicide attemps, using PQ are still frequent. Because <strong>of</strong> <strong>the</strong> well<br />

described pulmonary adverse effects, <strong>the</strong> use <strong>of</strong> PQ has been restricted in many<br />

countries, and rigorous tolerance limits on foods have been established.<br />

Beyond its use as an herbicide and a poison for suicide attempts, PQ has become a<br />

model for pro-oxidant <strong>induced</strong> chemical toxicity. Moreover, <strong>the</strong> knowledge about <strong>the</strong><br />

mechanism(s) <strong>of</strong> PQ toxicity has contributed significantly to <strong>the</strong> concept <strong>of</strong> cell-specific<br />

toxicity and has given rise to <strong>the</strong> notion that <strong>the</strong> cellular accumulation <strong>of</strong> a toxic agent<br />

through an endogenous transport system may underlie <strong>the</strong> observed toxic effects. The<br />

pulmonary effects <strong>of</strong> PQ can be readily explained by <strong>the</strong> participation <strong>of</strong> <strong>the</strong> PUS in its<br />

accumulation, a transporter, abundantly expressed in <strong>the</strong> membrane <strong>of</strong> alveolar cells<br />

type I and II and Clara cells. Fur<strong>the</strong>r downstream at <strong>the</strong> toxicodynamic level, <strong>the</strong> main<br />

pathways responsible for molecular mechanism <strong>of</strong> PQ toxicity are based on its redox<br />

cycling, with a constant flow <strong>of</strong> electrons to O2, with <strong>the</strong> consequent intracellular ROS<br />

generation, depletion <strong>of</strong> NADPH, formation <strong>of</strong> disulfides, protein oxidation, DNA<br />

damage, LPO and, in some cases, ensuing inflammatory reaction with formation <strong>of</strong><br />

massive fibrosis.<br />

More than 44 years after <strong>the</strong> first reports <strong>of</strong> PQ human poisonings, recovery in<br />

such cases remains poor and accepted treatment regimens are virtually nonexistent. As<br />

consequence <strong>of</strong> <strong>the</strong> rapid onset <strong>of</strong> <strong>the</strong> pulmonary injuries, treatments to decrease PQ<br />

absorption, extracorporeal elimination methods and <strong>the</strong> majority <strong>of</strong> <strong>the</strong> treatments<br />

related to pathophysiological lesions, are <strong>of</strong>ten inefficacious in modifying <strong>the</strong> clinical<br />

course. Despite <strong>the</strong> intensive <strong>research</strong> on PQ toxicity, nei<strong>the</strong>r <strong>the</strong> final cytotoxic<br />

mechanism nor a clinically useful antidote has yet been disclosed. Given <strong>the</strong><br />

considerable toxicity <strong>of</strong> PQ, is has been considered that treatments should be performed<br />

in groups <strong>of</strong> patients with a probability <strong>of</strong> survival over 20%, when employing <strong>the</strong><br />

nomogram described by Hart et al. (Hart et al., 1984). Treatments that may radically<br />

209


Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies________________________________________<br />

improve <strong>the</strong> prognosis should be able to remove PQ from <strong>the</strong> lung rapidly or should<br />

interrupt <strong>the</strong> toxic pathway before irreversible pulmonary cellular damage has occurred.<br />

210<br />

Bearing in mind <strong>the</strong> above-mentioned considerations, <strong>the</strong> global aims <strong>of</strong> this<br />

dissertation were to study <strong>the</strong> <strong>mechanisms</strong> <strong>of</strong> PQ-<strong>induced</strong> toxicity with special focus on<br />

its target <strong>organ</strong> – <strong>the</strong> lung - and to develop efficient antidotes to be used in human PQ<br />

poisonings. It is expected that an enhancement <strong>of</strong> <strong>the</strong> knowledge in this field, resulting<br />

from this dissertation, will provide medical doctors new tools for <strong>the</strong> difficult task <strong>of</strong><br />

treating PQ <strong>into</strong>xicated patients and thus to reduce <strong>the</strong> morbidity and mortality<br />

associated to this herbicide.<br />

During <strong>the</strong> last years, our <strong>research</strong> group has been a reference in <strong>the</strong> field <strong>of</strong> PQ<br />

toxicity to hospitals in <strong>the</strong> centre and north <strong>of</strong> Portugal. Accordingly, this dissertation<br />

was enriched by a description <strong>of</strong> a successful clinical case, regarding <strong>the</strong> <strong>into</strong>xication <strong>of</strong><br />

a 15-year-old girl by a presumed lethal dose <strong>of</strong> PQ, in which we participated actively<br />

through a deep discussion about <strong>the</strong> possible <strong>the</strong>rapeutic measures with <strong>the</strong> medical<br />

staff involved in <strong>the</strong> treatment, and later, in writing <strong>the</strong> manuscript (CHAPTER I).<br />

Besides <strong>the</strong> measures for decreasing PQ absorption and increasing its elimination from<br />

<strong>the</strong> blood, o<strong>the</strong>r protective procedures were applied aiming to reduce <strong>the</strong> production <strong>of</strong><br />

ROS, scavenge and repair ROS-<strong>induced</strong> lesions, and to reduce inflammation. The<br />

status-<strong>of</strong>-<strong>the</strong>-art concerning <strong>the</strong> biochemical and toxicological aspects <strong>of</strong> PQ poisoning<br />

and <strong>the</strong> pharmacological basis <strong>of</strong> <strong>the</strong> respective treatment protocol was presented. It was<br />

concluded that <strong>the</strong> intensive and aggressive treatment followed (CHP, pulse <strong>the</strong>rapy<br />

with CP and MP, vitamin-E, DFO and NAC), once high urinary and/or plasmatic PQ<br />

concentrations were detected, could constitute a promising treatment <strong>of</strong> PQ human<br />

<strong>into</strong>xications (see CHAPTER I for treatment protocol details). Besides <strong>the</strong> treatment, in<br />

this particular case, o<strong>the</strong>r good prognostic factors were <strong>the</strong> young age <strong>of</strong> <strong>the</strong> patient,<br />

lesser degrees <strong>of</strong> leukocytosis and acidosis, and <strong>the</strong> absence <strong>of</strong> renal, hepatic, and<br />

pancreatic failures on admission after acute PQ poisoning.<br />

In <strong>the</strong> CHAPTER II, <strong>the</strong> usefulness <strong>of</strong> <strong>the</strong> isolated rat lung model was explored to<br />

characterize <strong>the</strong> toxicokinetic behaviour <strong>of</strong> PQ in this tissue after bolus injection under<br />

standard experimental conditions and to evaluate <strong>the</strong> influence <strong>of</strong> iso-osmotic<br />

replacement <strong>of</strong> Na + by Li + in <strong>the</strong> perfusion medium. The obtained results showed that<br />

<strong>the</strong> isolated rat lung model is a very useful experimental procedure for PQ toxicokinetic<br />

analysis. It was also observed that Na + -depletion in <strong>the</strong> perfusion medium leads to a<br />

decreased uptake <strong>of</strong> PQ in <strong>the</strong> isolated rat lung, although it seems that this condition


________________________________________Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies<br />

does not contribute to improve <strong>the</strong> elimination <strong>of</strong> PQ once <strong>the</strong> herbicide reaches <strong>the</strong><br />

extravascular structures <strong>of</strong> <strong>the</strong> lung.<br />

Techniques <strong>of</strong> tissue isolation and perfusion <strong>of</strong>fer an excellent alternative to<br />

characterize <strong>the</strong> kinetic pr<strong>of</strong>ile for a tissue in a single animal, avoiding <strong>the</strong> inter-<br />

individual variability, and leading as well to a corresponding reduction in curve<br />

replicates and hence to a substantial reduction in <strong>the</strong> number <strong>of</strong> animals used (5-8<br />

versus 50-80/tissue). Never<strong>the</strong>less, PQ toxicity results from a myriad <strong>of</strong> factors,<br />

toge<strong>the</strong>r contributing to a death outcome and that required fur<strong>the</strong>r studies using in vivo<br />

approaches. Accordingly, <strong>the</strong> subsequent objective <strong>of</strong> this dissertation was to provide an<br />

effective solution to reduce <strong>the</strong> levels <strong>of</strong> PQ in <strong>the</strong> lung and, by this way, its toxicity<br />

(CHAPTER III). This approach is expected to increase <strong>the</strong> success <strong>of</strong> <strong>the</strong> treatments and<br />

consequently <strong>the</strong> survival <strong>of</strong> <strong>the</strong> PQ-<strong>into</strong>xicated patients. For that purpose, we evaluated<br />

<strong>the</strong> putative usefulness <strong>of</strong> <strong>the</strong> well known multidrug resistance (MDR) phenomena for<br />

clearing up lung PQ. MDR is characterized by <strong>the</strong> occurrence <strong>of</strong> cross-resistance <strong>of</strong><br />

cells to a broad range <strong>of</strong> structurally and functionally unrelated xenobiotics (Gottesman<br />

and Pastan, 1993). Several <strong>mechanisms</strong> are involved in MDR. One <strong>of</strong> <strong>the</strong> most wellknown<br />

<strong>mechanisms</strong> is <strong>the</strong> overexpression <strong>of</strong> a plasma membrane phosphoglycoprotein<br />

termed P-glycoprotein (P-gp). P-gp, a member <strong>of</strong> ATP-binding cassette (ABC)<br />

transporter superfamily, was initially identified in tumour cells as an ATP-dependent<br />

transporter, which can export a wide variety <strong>of</strong> unmodified substrates out <strong>of</strong> <strong>the</strong> cell<br />

(Ling et al., 1983; Chen et al., 1986 ; Cordon-Cardo et al., 1990; Gottesman and Pastan,<br />

1993). Besides tumour cells, P-gp was also found to be expressed in a polarized manner,<br />

at <strong>the</strong> apical surface (or luminal, depending on <strong>the</strong> <strong>organ</strong>) in a variety <strong>of</strong> normal tissues,<br />

including <strong>the</strong> lungs (Crapo et al., 1982). Such a spatial distribution <strong>of</strong> this efflux<br />

transporter represents a functional important element in reducing <strong>the</strong> systemic exposure<br />

and specific tissue access <strong>of</strong> potentially harmful xenobiotics. The expression <strong>of</strong> P-gp in<br />

liver, brain, and intestinal tissue and also in lung tissue has been shown to be <strong>induced</strong> by<br />

DEX (Demeule et al., 1999). This increased expression is rapid, since it is observed to<br />

be maximal after only one day post-treatment (Demeule et al., 1999). In CHAPTER III<br />

it was demonstrated that <strong>the</strong> induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-gp by DEX (100<br />

mg/Kg i.p.), two hours after administration <strong>of</strong> a lethal dose <strong>of</strong> PQ (25 mg/Kg i.p.) to<br />

Wistar rats, results in a remarkable decrease <strong>of</strong> lung PQ levels (to about 40% <strong>of</strong> <strong>the</strong> only<br />

PQ-exposed group in just 24 hours) and an increase <strong>of</strong> its faecal excretion. As expected,<br />

<strong>the</strong> decrease <strong>of</strong> lung PQ levels resulted in <strong>the</strong> prevention <strong>of</strong> PQ-<strong>induced</strong> lung toxicity,<br />

211


Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies________________________________________<br />

which was evidenced by a significant decrease <strong>of</strong> several biochemical and<br />

histopathological biomarkers <strong>of</strong> toxicity. Verapamil [VER (10 mg/kg i.p.)], a<br />

competitive inhibitor <strong>of</strong> P-gp, given one hour before DEX, blocked its protective<br />

effects, and lead to an increase <strong>of</strong> PQ lung concentration (up to about twice <strong>of</strong> <strong>the</strong> only<br />

PQ-exposed group in just 24 hours) and toxicity, indicating <strong>the</strong> important role <strong>of</strong> this<br />

transporter in PQ excretion. The obtained results showed that DEX also ameliorated <strong>the</strong><br />

biochemical and histological liver alterations <strong>induced</strong> by PQ in Wistar rats (CHAPTER<br />

IV). On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong>se improvements were not observed in kidney and spleen <strong>of</strong><br />

DEX treated rats. Notwithstanding <strong>the</strong> conflicting findings, <strong>the</strong> sum <strong>of</strong> <strong>the</strong>se effects was<br />

clearly positive, since it was observed an increased survival rate to 50% 10 days post<strong>into</strong>xication,<br />

which indicates that high dosage DEX treatment constitutes an important<br />

and valuable <strong>the</strong>rapeutic tool to be used against PQ-<strong>induced</strong> toxicity. This approach is<br />

still to be applied in human PQ poisonings, but constitutes a landmark in <strong>the</strong> fight<br />

against PQ-<strong>induced</strong> toxicity, considering that this is <strong>the</strong> first time that an accelerated<br />

release <strong>of</strong> PQ taken up by <strong>the</strong> lungs is achieved.<br />

According to <strong>the</strong> promising results in <strong>the</strong> treatment <strong>of</strong> PQ toxicity, in <strong>the</strong> scope <strong>of</strong><br />

this dissertation it was patented <strong>the</strong> process <strong>of</strong> induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-gp for<br />

<strong>the</strong> treatment <strong>of</strong> xenobiotic-<strong>induced</strong> <strong>into</strong>xications in mammals, assuming that <strong>the</strong><br />

successful delivery <strong>of</strong> <strong>the</strong> inducer to <strong>the</strong> target tissue is possible. In fact, <strong>the</strong> subject <strong>of</strong><br />

<strong>the</strong> patent claims is precisely <strong>the</strong> opposite <strong>of</strong> <strong>the</strong> anticancer <strong>the</strong>rapy, in which <strong>the</strong> main<br />

objective is to limit <strong>the</strong> drug efflux <strong>of</strong> <strong>the</strong> cells.<br />

212<br />

Successful realization <strong>of</strong> <strong>the</strong> value proposition is contingent upon <strong>the</strong> following:<br />

Targeting <strong>the</strong> inducers to <strong>the</strong> appropriate tissue without increasing expression in<br />

o<strong>the</strong>r tissues, where increased xenobiotic resistance is not desired;<br />

Low toxicity pr<strong>of</strong>ile and acceptable <strong>the</strong>rapeutic index;<br />

Research on drug-drug interactions for safety purposes, especially when used as<br />

a co-<strong>the</strong>rapeutic;<br />

Any o<strong>the</strong>r considerations associated with regulatory approvals.<br />

If adequate delivery to <strong>the</strong> target tissue is achieved and <strong>the</strong> inducer <strong>of</strong> de novo<br />

syn<strong>the</strong>sis <strong>of</strong> P-gp does not interfere with <strong>the</strong> activity <strong>of</strong> <strong>the</strong> <strong>the</strong>rapeutic agent <strong>of</strong> interest


________________________________________Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies<br />

as a co-<strong>the</strong>rapeutic, or create or exacerbate side effects, and <strong>the</strong> inducer reduces <strong>organ</strong><br />

toxicity attributed to <strong>the</strong> agent, <strong>the</strong> applicative potential is manifest. In addition,<br />

prevention <strong>of</strong> disease by reducing exposure <strong>of</strong> target tissue to risk factors will be <strong>of</strong><br />

interest.<br />

In spite <strong>of</strong> our proposed antidotal pathway, through <strong>the</strong> induction <strong>of</strong> de novo<br />

syn<strong>the</strong>sis <strong>of</strong> P-gp, <strong>the</strong> persistent lacuna related to <strong>the</strong> inexistence <strong>of</strong> an antidote that<br />

conducts to 100% <strong>of</strong> survival, impelled <strong>the</strong> work described in <strong>the</strong> CHAPTER V. This<br />

study concerns to <strong>the</strong> use <strong>of</strong> sodium salicylate (NaSAL) in <strong>the</strong> treatment <strong>of</strong> <strong>into</strong>xications<br />

caused by <strong>the</strong> PQ. The role <strong>of</strong> <strong>the</strong> oxidative stress, platelet aggregation, nuclear factor<br />

(NF)-κB activation and fibrosis in PQ-<strong>induced</strong> lung toxicity, as well as <strong>the</strong> remarkable<br />

healing effects obtained by <strong>the</strong> administration <strong>of</strong> sodium salicylate (NaSAL, 200 mg/Kg<br />

i.p.), were assessed. The results clearly showed that NaSAL has a great potential to be<br />

used as an antidote against PQ-<strong>induced</strong> lung toxicity, mainly mediated by an effective<br />

inhibition <strong>of</strong> pro-inflammatory factors such as NF-κB, by scavenging ROS, and also<br />

through <strong>the</strong> inhibition <strong>of</strong> myeloperoxidase activity and inhibition <strong>of</strong> platelet aggregation<br />

(Fig. 19). The obtained results exceeded our best expectations since not only <strong>the</strong> toxicity<br />

was reverted but, most significantly, it was observed <strong>the</strong> full survival <strong>of</strong> <strong>the</strong> PQ-<br />

<strong>into</strong>xicated rats treated with NaSAL (extended for more than 30 days) in opposition to<br />

100% <strong>of</strong> mortality by <strong>the</strong> day 6 in PQ-only exposed animals. NaSAL seems <strong>the</strong>n to<br />

constitute a real antidote for PQ poisonings, since it is <strong>the</strong> first compound with such<br />

degree <strong>of</strong> success. With this study, it was given an important step in <strong>the</strong> treatment <strong>of</strong> PQ<br />

<strong>into</strong>xications. Due to <strong>the</strong> importance <strong>of</strong> this study, <strong>the</strong> use <strong>of</strong> salicylates and derivatives<br />

as antidotes in PQ <strong>into</strong>xication were patented. One may consider that <strong>the</strong> dose <strong>of</strong><br />

NaSAL used in this study is quite high. Accordingly to literature, <strong>the</strong> pathophysiologic<br />

changes attributable to high doses <strong>of</strong> NaSAL result in various clinical manifestations<br />

depending on <strong>the</strong> amount ingested; in humans, an oral dose <strong>of</strong> 200 mg/Kg yields,<br />

approximately, a serum concentration <strong>of</strong> 500-770 mg/L (Proudfoot, 1983; Yip et al.,<br />

1994; Dargan et al., 2002). These serum levels originate signals <strong>of</strong> mild side effects<br />

such as nausea, vomiting, tinnitus, hyperventilation and respiratory alkalosis.<br />

Never<strong>the</strong>less, in <strong>the</strong> particular case <strong>of</strong> life-threatening PQ poisonings, <strong>the</strong> risk/benefit<br />

ratio will most probably flip to <strong>the</strong> beneficial effects <strong>of</strong> NaSAL, although this still needs<br />

to be carefully confirmed in clinical trials.<br />

Of note, <strong>the</strong> administrations <strong>of</strong> DEX and NaSAL were given two hours after<br />

<strong>into</strong>xication <strong>of</strong> rats with PQ, a lag time that confers realism to be applied in humans,<br />

213


Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies________________________________________<br />

since this chronological time corresponds to longer biological time in humans and<br />

<strong>the</strong>refore it may represent <strong>the</strong> actual time that passes between <strong>the</strong> herbicide ingestion<br />

and <strong>the</strong> begin <strong>of</strong> <strong>the</strong> treatments.<br />

214<br />

Finally in <strong>the</strong> CHAPTER VI <strong>of</strong> this dissertation, it was studied <strong>the</strong> occurrence <strong>of</strong><br />

apoptotic events in <strong>the</strong> lung <strong>of</strong> male Wistar rats, 24, 48 and 96 hours after PQ-exposure<br />

(25 mg/Kg i.p.) as well as <strong>the</strong> putative healing effects provided by NaSAL (200 mg/Kg<br />

i.p.), when administered two hours after PQ. PQ-exposure resulted in marked lung<br />

apoptosis, in a time dependent manner, characterized by <strong>the</strong> “ladder-like” pattern <strong>of</strong><br />

DNA observed through electrophoresis and by <strong>the</strong> presence <strong>of</strong> terminal<br />

deoxynucleotidyl transferase-mediated deoxyuridine triphosphate nick end-labeling<br />

(TUNEL)-positive cells (TPC) as revealed by immunohistochemistry. PQ-exposed rats<br />

suffered a time-dependent increase <strong>of</strong> caspase-3 and caspase-8 and a decrease <strong>of</strong><br />

caspase-1 activities in <strong>the</strong> lung. Also observed, was a marked mitochondrial<br />

dysfunction, evidenced by cytochrome c (Cyt c) release, and a transcriptional activation<br />

<strong>of</strong> <strong>the</strong> p53 and activator protein-1 (AP-1) transcription factors, in a time-dependent<br />

manner as a consequence <strong>of</strong> PQ-exposure. Overall, this work led to a better<br />

understanding about <strong>the</strong> <strong>mechanisms</strong> related to <strong>the</strong> PQ-<strong>induced</strong> toxicity in <strong>the</strong><br />

respiratory tract, showing that PQ induces several events involved in <strong>the</strong> apoptotic<br />

pathways, which might trigger its lung toxicity. The data reported in <strong>the</strong> CHAPTER VI<br />

also reinforced <strong>the</strong> potential use <strong>of</strong> NaSAL in <strong>the</strong> protection against PQ-<strong>induced</strong> lung<br />

damage. NaSAL treatment resulted in <strong>the</strong> remission <strong>of</strong> <strong>the</strong> observed apoptotic signaling<br />

and consequently <strong>of</strong> lung apoptosis. Thus, it is legitimate to speculate that <strong>the</strong> strong<br />

protection (which ended up in full survival) conferred by NaSAL to PQ-<strong>into</strong>xicated rats<br />

(CHAPTER V), besides inhibition <strong>of</strong> NF-κB activation, MPO activity and platelet<br />

aggregation, and scavenging <strong>of</strong> ROS, also results from <strong>the</strong> blockade <strong>of</strong> <strong>the</strong> intrinsic and<br />

extrinsic apoptotic pathways (Fig. 17).


________________________________________Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies<br />

↑ MPO<br />

activity<br />

Platelet<br />

Aggregation<br />

Pulmonary<br />

Edema<br />

Fig. 17 – Proposed protective <strong>mechanisms</strong> <strong>of</strong> sodium salicylate against pulmonary<br />

<strong>paraquat</strong> toxicity.<br />

NF-kB<br />

Activation<br />

Type I, II and<br />

Clara cells<br />

disruption<br />

x<br />

x<br />

x<br />

x<br />

x<br />

Caspase<br />

cascade<br />

Lipid<br />

peroxidation<br />

AP-1 and p53<br />

activation<br />

Oxidative<br />

Stress<br />

x x<br />

x<br />

x x<br />

x<br />

x<br />

x<br />

x x<br />

x<br />

x x<br />

x<br />

x<br />

x<br />

PARAQUAT SODIUM<br />

SALICYLATE<br />

SODIUM SALICYLATE x<br />

Collagen<br />

Deposition<br />

DNA<br />

Fragmentation<br />

Protein<br />

oxidation<br />

Infiltration <strong>of</strong><br />

inflammatory<br />

cells<br />

The results obtained in <strong>the</strong> ambit <strong>of</strong> this dissertation suggest that <strong>the</strong>se two last<br />

<strong>the</strong>rapeutic approaches have high potential to be applied in humans. Although <strong>the</strong> dose<br />

<strong>of</strong> DEX to induce de novo syn<strong>the</strong>sis <strong>of</strong> P-gp in lungs and <strong>the</strong> dose <strong>of</strong> NaSAL necessary<br />

to inhibit NF-κB activation and related effects are quite high, and mild toxicity may<br />

ensue following <strong>the</strong> administration <strong>of</strong> such high doses in humans, clinically, PQ<br />

poisoning is an extremely frustrating condition to manage, due to <strong>the</strong> elevated morbidity<br />

and mortality observed so far, which may endorse this type <strong>of</strong> drastic treatment.<br />

C<br />

C<br />

C O<br />

H<br />

C<br />

215<br />

C O


Part III – Integrated overview <strong>of</strong> <strong>the</strong> performed studies________________________________________<br />

216


_________________________________________________________________Part III – Conclusions<br />

PART III<br />

2. CONCLUSIONS<br />

2. CONCLUSIONS<br />

217


Part III – Conclusions_________________________________________________________________<br />

218


_________________________________________________________________Part III – Conclusions<br />

2. CONCLUSIONS<br />

I. The <strong>the</strong>rapeutic protocol (CHP, pulse <strong>the</strong>rapy with CP and MP, vitamin-E, DFO<br />

and NAC) followed in <strong>the</strong> reported clinical case resulted in a positive outcome,<br />

which reinforces its potential use in <strong>the</strong> treatment <strong>of</strong> PQ human <strong>into</strong>xications;<br />

II. The isolated rat lung model is useful for <strong>the</strong> study <strong>of</strong> PQ toxicokinetics;<br />

III. The polyexponential pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> efferent fluid curves, in <strong>the</strong> isolated rat lung<br />

model, revealed a rapid access <strong>of</strong> PQ to extravascular spaces with a slow washout<br />

process;<br />

IV. Lung uptake <strong>of</strong> PQ was Na + dependent. This condition did not contribute to<br />

improve <strong>the</strong> elimination <strong>of</strong> PQ once <strong>the</strong> herbicide reached <strong>the</strong> extravascular<br />

structures <strong>of</strong> <strong>the</strong> lung;<br />

V. The induction <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-gp by DEX (100 mg/Kg i.p.), two hours<br />

after PQ exposure (25 mg/Kg i.p.), resulted in a remarkable decrease <strong>of</strong> PQ lung<br />

accumulation and an increase <strong>of</strong> its faecal excretion, in Wistar rats;<br />

VI. The decrease <strong>of</strong> lung PQ levels, resulting from <strong>the</strong> induction <strong>of</strong> de novo syn<strong>the</strong>sis<br />

<strong>of</strong> P-gp by DEX, led to a remarkable amelioration <strong>of</strong> practically all toxicological<br />

parameters that were changed by PQ exposure. This treatment aggravated <strong>the</strong> PQ<strong>induced</strong><br />

toxicity in <strong>the</strong> kidneys and spleen;<br />

VII. The apparent protection that high dosage DEX treatment awards to <strong>the</strong> lungs and<br />

liver <strong>of</strong> <strong>the</strong> PQ-<strong>into</strong>xicated animals outweighed <strong>the</strong> increased damage to <strong>the</strong>ir<br />

spleens and kidneys, reflected by a higher survival rate, indicating that DEX<br />

treatment may constitute an important and valuable <strong>the</strong>rapeutic drug to be used<br />

against PQ-<strong>induced</strong> toxicity;<br />

219


Part III – Conclusions_________________________________________________________________<br />

VIII. VER (10 mg/Kg i.p.), a competitive inhibitor <strong>of</strong> P-gp, given one hour before<br />

220<br />

DEX, blocked its protective effects, and led to an increase <strong>of</strong> PQ lung<br />

concentration (up to about twice <strong>of</strong> <strong>the</strong> only PQ-exposed group in just 24 hours)<br />

and toxicity, indicating <strong>the</strong> important role <strong>of</strong> this transporter in PQ excretion;<br />

IX. NaSAL has a great potential to be used as an antidote against PQ poisonings,<br />

mainly mediated by counteracting <strong>the</strong> following PQ-<strong>induced</strong> toxic effects:<br />

- Increased activation <strong>of</strong> pro-inflammatory factors, specifically NF-κB;<br />

- Increased formation <strong>of</strong> ROS;<br />

- Increased myeloperoxidase activity;<br />

- Increase <strong>of</strong> platelet aggregation;<br />

- Time-dependent increase <strong>of</strong> caspase-3 and caspase-8 and a decrease <strong>of</strong> caspase-1<br />

activities;<br />

- Marked mitochondrial dysfunction evidenced by Cyt c release and <strong>the</strong> transcriptional<br />

activation <strong>of</strong> <strong>the</strong> p53 and AP-1 transcription factors;<br />

- Increased apoptosis;<br />

X. Administration <strong>of</strong> NaSAL to Wistar rats (200 mg/kg i.p.), two hours after<br />

exposure to a toxic dose <strong>of</strong> PQ (25 mg/kg, i.p.) resulted in full survival <strong>of</strong> <strong>the</strong> PQtreated<br />

rats (extended for more than 30 days) in comparison with 100% <strong>of</strong><br />

mortality by day 6 in animals exposed only to PQ. NaSAL constitutes <strong>the</strong> first<br />

compound with such degree <strong>of</strong> success (100% survival).


__________________________________________________Part III – Directions for future <strong>research</strong>es<br />

PART III<br />

3. DIRECTION FOR FUTURE RESEARCHES<br />

3. DIRECTIONS FOR FUTURE RESEARCH<br />

221


Part III – Directions for future <strong>research</strong>es__________________________________________________<br />

222


__________________________________________________Part III – Directions for future <strong>research</strong>es<br />

3. DIRECTIONS FOR FUTURE RESEARCH<br />

Future projects are required to continue with <strong>the</strong> pre-clinical studies to explain, in<br />

more detail, <strong>the</strong> mode <strong>of</strong> action <strong>of</strong> salicylates in <strong>the</strong> protection against PQ-<strong>induced</strong> lung<br />

damage, but also pre-clinical studies, particularly those aimed to syn<strong>the</strong>size new,<br />

specific and more potent inducers <strong>of</strong> de novo syn<strong>the</strong>sis <strong>of</strong> P-gp in attempt to avoid <strong>the</strong><br />

well known corticosteroid adverse effects. In addition, <strong>the</strong>se P-gp de novo syn<strong>the</strong>sis<br />

inducers may reveal <strong>the</strong>irselves to be important drugs to reduce systemic exposure and<br />

specific tissue access <strong>of</strong> several potential harmful xenobiotics. Finally, <strong>the</strong>se drugs will<br />

increase <strong>the</strong> <strong>the</strong>rapeutic arsenal available, giving more confidence and security to <strong>the</strong><br />

physicians when administering drugs with narrow <strong>the</strong>rapeutic windows. P-gp inducers<br />

have an obvious potential wide application and until now no similar drug has yet been<br />

developed. In addition, more studies are required to evaluate <strong>the</strong> beneficial effects <strong>of</strong><br />

<strong>the</strong>se <strong>the</strong>rapeutic approaches using <strong>multiple</strong> administrations <strong>of</strong> <strong>the</strong>se or o<strong>the</strong>r drugs that<br />

may reduce <strong>the</strong> respective side effects <strong>of</strong> high doses and thus to mimic what is done at<br />

<strong>the</strong> hospital critical care departments leading with this very common <strong>into</strong>xication. It<br />

should be also objective <strong>of</strong> future projects, to carry out studies using oral route for PQ<br />

administration, since ingestion is <strong>the</strong> most common way <strong>of</strong> <strong>into</strong>xication, and our<br />

previous studies were done intraperitoneally. Forensic studies in post-mortem samples<br />

from human <strong>into</strong>xications should be also performed. At <strong>the</strong> end, <strong>the</strong> aim is to apply <strong>the</strong><br />

developed antidotes in humans <strong>into</strong>xicated by PQ.<br />

223


Part III – Directions for future <strong>research</strong>es__________________________________________________<br />

224


________________________________________________________Part IV – References<br />

1. REFERENCES<br />

PART IV<br />

1. REFERENCES<br />

225


Part IV – References____________________________________________________________________<br />

226


________________________________________________________Part IV – References<br />

Ackrill, P., Hasleton, P. S., and Ralston, A. J. (1978). Oesophageal perforation due to<br />

<strong>paraquat</strong>. Br Med J 1, 1252-1253.<br />

Addo, E., and Poon-King, T. (1986). Leukocyte suppression in treatment <strong>of</strong> 72 patients<br />

with <strong>paraquat</strong> poisoning. Lancet 1, 1117-1120.<br />

Addo, E., Ramdial, S., and Poon-King, T. (1984). High dosage cyclophosphamide and<br />

dexamethasone treatment <strong>of</strong> <strong>paraquat</strong> poisoning with 75% survival. West Indian<br />

Med J 33, 220-226.<br />

Akhavein, A. A., and Linscott, D. L. (1968). The dipyridylium herbicides, <strong>paraquat</strong> and<br />

diquat. Residue Rev 23, 97-145.<br />

Aleksic-Kovacevic, S., Knezevic, M., Jovanovic, M., and Jelesijevic, T. (2003).<br />

Accidental poisoning <strong>of</strong> dogs with herbicide <strong>paraquat</strong>. J Vet Pharmacol Ther 26<br />

(Suppl. 1), 268.<br />

Ali, S., Jain, S. K., Abdulla, M., and Athar, M. (1996). Paraquat <strong>induced</strong> DNA damage<br />

by reactive oxygen species. Biochem Mol Biol Int 39, 63- 67.<br />

Almog, C., and Tal, E. (1967). Death from <strong>paraquat</strong> after subcutaneous injection. Br<br />

Med J 3, 721.<br />

Amondham, W., Parkpian, P., Polprasert, C., DeLaune, R. D., and Jugsujinda, A.<br />

(2006). Paraquat adsorption, degradation, and remobilization in tropical soils <strong>of</strong><br />

Thailand. J Environ Sci Health B 41, 485-507.<br />

Anderson, M. E. (1997). Glutathione and glutathione delivery compounds. Adv<br />

Pharmacol 38, 65-78.<br />

Aoki, H., Otaka, Y., Igarashi, K., and Takenaka, A. (2002). Soy protein reduces<br />

<strong>paraquat</strong>-<strong>induced</strong> oxidative stress in rats. J Nutr 132, 2258-2262.<br />

Ariyama, J., Shimada, H., Aono, M., Tsuchida, H., and Hirai, K. I. (2000). Prop<strong>of</strong>ol<br />

improves recovery from <strong>paraquat</strong> acute toxicity in vitro and in vivo. Intensive<br />

Care Med 26, 981-987.<br />

Athanaselis, S., Qammaz, S., Alevisopoulos, G., and Koutselinis, A. (1983).<br />

Percutaneous <strong>paraquat</strong> <strong>into</strong>xication. J Toxicol Cut Ocular Toxicol 2, 3-5.<br />

Autor, A. P. (1977). Biochemical Mechanisms <strong>of</strong> Paraquat Toxicity. Academic Press,<br />

New York.<br />

Aziz, S. M., Lipke, D. W., Olson, J. W., and Gillespie, M. N. (1994). Role <strong>of</strong> ATP and<br />

sodium in polyamine transport in bovine pulmonary artery smooth cells.<br />

Biochem Pharmacol 48, 1611-1618.<br />

Bagchi, D., Prasad, R., and Das, D. K. (1989). Direct scavenging <strong>of</strong> free radicals by<br />

captopril, an angiotensin-converting enzyme inhibitor. Biochem Biophys Res<br />

Commun 158, 52-57.<br />

Baldwin, B. C., Bray, M. F., and Geoghegan, M. J. (1966). The microbial<br />

decomposition <strong>of</strong> <strong>paraquat</strong>. Biochem J 101, 15.<br />

Baselt, R. C., and Cravey, R. H. (1989). Paraquat. In Disposition <strong>of</strong> Toxic Drugs and<br />

Chemicals in Man (R. C. Baselt, and R. H. Cravey, Eds.), pp. 637-640. Year<br />

Book, Chicago.<br />

Bassett, D. J., and Fisher, A. B. (1978). Alterations <strong>of</strong> glucose metabolism during<br />

perfusion <strong>of</strong> rat lung with <strong>paraquat</strong>. Am J Physiol 234, E653-E659.<br />

Bataller, R., Bragulat, E., Nogue, S., Gorbig, M. N., Bruguera, M., and Rodes, J. (2000).<br />

Prolonged cholestasis after acute <strong>paraquat</strong> poisoning through skin absorption.<br />

Am J Gastroenterol 95, 1340-1343.<br />

Bateman, D. N. (1987). Pharmacological treatments <strong>of</strong> <strong>paraquat</strong> poisoning. Hum<br />

Toxicol 6, 57–62.<br />

227


Part IV – References____________________________________________________________________<br />

Bauman, J. W., Liu, J., Liu, Y. P., and Klaassen, C. D. (1991). Increase in<br />

metallothionein produced by chemicals that induce oxidative stress. Toxicol<br />

Appl Pharmacol 110, 347-354.<br />

Beebeejaun, A. R., Beevers, G., and Rogers, W. N. (1971). Paraquat poisoningprolonged<br />

excretion. Clin Toxicol 4, 397-407.<br />

Behne, D., and Wolters, W. (1983). Distribution <strong>of</strong> selenium and glutathione peroxidase<br />

in <strong>the</strong> rat. J Nutr 113, 456-461.<br />

Berisha, H. I., Pakbaz, H., Absood, A., and Said, S. I. (1994). Nitric oxide as a mediator<br />

<strong>of</strong> oxidant lung injury due to <strong>paraquat</strong>. Proc Natl Acad Sci U S A 91, 7445-7449.<br />

Berry, D. J., and Grove, J. (1971). The determination <strong>of</strong> <strong>paraquat</strong> (I,I'-dimethyl-4,4'bipyridylium<br />

cation) in urine. Clin Chim Acta 34, 5-11.<br />

Bier, R. K., and Osborne, I. J. (1978). Pulmonary changes in <strong>paraquat</strong> poisoning.<br />

Radiology 127, 308.<br />

Binns, C. W. (1976). A deadly cure for lice - a case <strong>of</strong> <strong>paraquat</strong> poisoning. P N G Med J<br />

19, 105-107.<br />

Bird, C. L., and Kuhn, A. T. (1981). Electrochemistry <strong>of</strong> <strong>the</strong> Viologens. Chem Soc Rev<br />

10, 49-82.<br />

Bismuth, C., Baud, F. J., Garnier, R., Muszinski, J., and Houze, P. (1988). Paraquat<br />

poisoning: biological presentation. J Toxicol Clin Exp 8, 211-218.<br />

Bismuth, C., Garnier, R., Dally, S., Fournier, P. E., and Scherrmann, J. M. (1982).<br />

Prognosis and treatment <strong>of</strong> <strong>paraquat</strong> poisoning. A review <strong>of</strong> 28 cases. J Toxicol<br />

Clin Toxicol 19, 461–474.<br />

Bismuth, C., Hall, A., and Wong, A. (1995). Paraquat ingestion exposure:<br />

symptomatology and risk. In Paraquat poisoning: <strong>mechanisms</strong>, prevention,<br />

treatment (C. Bismuth, and A. H. Hall, Eds.), pp. 195-210. Marcel Dekker, New<br />

York.<br />

Bismuth, C., and Hall, A. H. (1995). Paraquat Poisoning: Mechanisms, Prevention,<br />

Treatment. Marcel Dekker, New York.<br />

Bismuth, C., Scherrmann, J. M., Garnier, R., Baud, F. J., and Pontal, P. G. (1987).<br />

Elimination <strong>of</strong> <strong>paraquat</strong>. Hum Toxicol 6, 63-67.<br />

Block, E. R. (1979). Potentiation <strong>of</strong> acute <strong>paraquat</strong> toxicity by vitamin E deficiency.<br />

Lung 156, 195-203.<br />

Bloodworth, L. L., Kershaw, J. B., Stevens, P. E., Alcock, C. J., and Rainford, D. J.<br />

(1986). Failure <strong>of</strong> radio<strong>the</strong>rapy to reverse progressive pulmonary fibrosis caused<br />

by <strong>paraquat</strong>. Br J Radiol 59, 1037-1039.<br />

Booz, G. W., Schorb, W., Dostal, D. E., Conrad, K. M., Chang, K. C., and Baker, K. M.<br />

(1993). Angiotensin II is a mitogenic in neonatal rat cardiac fibroblast. Circ Res<br />

72, 1245-1254.<br />

Boumpas, D. T., Austin, H. r., Vaughn, E. M., Klippel, J. H., Steinberg, A. D., Yarboro,<br />

C. H., and Balow, J. E. (1992). Controlled trial <strong>of</strong> pulse methylprednisolone<br />

versus two regimens <strong>of</strong> pulse cyclophosphamide in severe lupus nephritis.<br />

Lancet 340, 741-745.<br />

Bowles, M., Johnston, S. C., Scho<strong>of</strong>, D. D., Pentel, P. R., and Pond, S. M. (1988). Large<br />

scale production and purification <strong>of</strong> <strong>paraquat</strong> and desipramine monoclonal<br />

antibodies and <strong>the</strong>ir Fab fragments. Int J Immunopharmacol 10, 537-545.<br />

Bowles, M. R., Mulhern, T. D., Gordon, R. B., Inglis, H. R., Sharpe, I. A., Cogill, J. L.,<br />

and Pond, S. M. (1997). Bound Tris confounds <strong>the</strong> identification <strong>of</strong> binding site<br />

residues in a <strong>paraquat</strong> single chain antibody. J Biochem (Tokyo) 122.<br />

Braithwaite, R. A. (1987). Emergency analysis <strong>of</strong> <strong>paraquat</strong> in biological fluids. Hum<br />

Toxicol 6, 91-93.<br />

228


________________________________________________________Part IV – References<br />

Bramley, A., and Hart, T. B. (1983). Paraquat poisoning in <strong>the</strong> United Kingdom. Hum<br />

Toxicol 2, 417.<br />

Brigelius, R., Lenzen, R., and Sies, H. (1982). Increase in hepatic mixed disulphide and<br />

glutathione disulphide levels elicited by <strong>paraquat</strong>. Biochem Pharmacol 31, 1637-<br />

1641.<br />

Brooks, R. E. (1971). Ultrastructure <strong>of</strong> lung lesions produced by ingested chemicals. I.<br />

Effect <strong>of</strong> <strong>the</strong> herbicide <strong>paraquat</strong> on mouse lung. Lab Invest 25, 536-545<br />

Brown, O. R., Heitkamp, M., and Song, C. S. (Science). Niacin Reduces Paraquat<br />

Toxicity in Rats. 1981 212, 1510-1512.<br />

Buckley, N. A. (2001). Pulse corticosteroids and cyclophosphamide in <strong>paraquat</strong><br />

poisoning. Am J Respir Crit Care Med 163, 585.<br />

Buettner, G. R., and Jurkiewicz, B. A. (1996). Catalytic metals, ascorbate and free<br />

radicals. Combination to avoid. Rad Res 145, 532-541.<br />

Bullivant, C. M. (1966). Accidental poisoning by <strong>paraquat</strong>: Report <strong>of</strong> two cases in man.<br />

Br Med J 5498, 1272-1273.<br />

Burk, R. F., Lawrence, R. A., and Lane, J. M. (1980). Liver necrosis and lipid<br />

peroxidation in <strong>the</strong> rat as result <strong>of</strong> <strong>paraquat</strong> and diquat administration. Effect <strong>of</strong><br />

selenium deficiency. J Clin Invest 65, 1024-1031.<br />

Burns, R. G., and Audus, L. J. (1970). Distribution and breakdown <strong>of</strong> <strong>paraquat</strong> in soil.<br />

Weed Res 10, 49-58.<br />

Burton, G. W. (1994). Vitamin E: molecular and biological function. Proc Nutr Soc 53,<br />

251–262.<br />

Bus, J. S., Aust, S. D., and Gibson, J. E. (1974). Superoxide- and singlet oxygencatalyzed<br />

lipid peroxidation as a possible mechanism for <strong>paraquat</strong> (methyl<br />

viologen) toxicity. Biochem Biophys Res Commun 58, 749-755.<br />

Bus, J. S., Aust, S. D., and Gibson, J. E. (1975). Lipid peroxidation: a possible<br />

mechanism for <strong>paraquat</strong> toxicity. Res Commun Chem Pathol Pharmacol 11, 31-<br />

38.<br />

Bus, J. S., Cagen, S. Z., Olgaard, M., and Gibson, J. E. (1976). A mechanism <strong>of</strong><br />

<strong>paraquat</strong> toxicity in mice and rats. Toxicol. Appl. Pharmacol 35, 501-513.<br />

Butler, C., 2nd. (1975). Pulmonary interstitial fibrosis from <strong>paraquat</strong> in <strong>the</strong> hamster.<br />

Arch Pathol 99, 503-507.<br />

Butler, C. n., and Kleinerman, J. (1971). Paraquat in <strong>the</strong> rabbit. Br J Ind Med 28, 67-71.<br />

Cadot, R., Descotes, J., Grenot, C., Cuilleron, C. Y., and Evreux, J. C. (1985). Increased<br />

plasma <strong>paraquat</strong> levels in <strong>into</strong>xicated mice following anti<strong>paraquat</strong> F(ab')2<br />

treatment. J Immunopharmacol 7, 467-477.<br />

Cagen, S. Z., and Gibson, J. E. (1977). Liver damage following <strong>paraquat</strong> in seleniumdeficient<br />

and diethyl maleate-pretreated mice. Toxicol Appl Pharmacol 40, 193-<br />

200.<br />

Calderbank, A. (1968). The bipyridilium herbicides. Adv Pest Control Res 8, 129-235.<br />

Calderbank, A., and Slade, P. (1976). Diquat and <strong>paraquat</strong>. In Herbicide chemistry,<br />

degradation and mode <strong>of</strong> action (P. C. Kearne, and D. D. Kaufman, Eds.), pp.<br />

501-540. Dekker, New York.<br />

Campbell, S. (1968). Death from <strong>paraquat</strong> in a child. Lancet 1, 144.<br />

Candan, F., and Alagozlu, H. (2001). Captopril inhibits <strong>the</strong> pulmonary toxicity <strong>of</strong><br />

<strong>paraquat</strong> in rats. Hum Exp Toxicol 20, 637-641.<br />

Cant, J. S., and Lewis, D. R. (1968). Ocular damage due to <strong>paraquat</strong> and diquat. Br Med<br />

J 3, 59.<br />

229


Part IV – References____________________________________________________________________<br />

Cappelletti, G., Maggioni, M. G., and Maci, R. (1998). Apoptosis in human lung<br />

epi<strong>the</strong>lial cells: triggering by <strong>paraquat</strong> and modulation by antioxidants. Cell Biol<br />

Int 22, 671–678.<br />

Carr, A., and Frei, B. (1999). Does vitamin C act as a pro-oxidant under physiological<br />

conditions? FASEB J 13, 1007-1024.<br />

Carson, D. J., and Carson, E. D. (1976). The increasing use <strong>of</strong> <strong>paraquat</strong> as a suicidal<br />

agent. Forensic Sci 7, 151-160.<br />

Carson, E. D. (1972). Fatal <strong>paraquat</strong> poisoning in Nor<strong>the</strong>rn Ireland. J Forensic Sci Soc<br />

12, 437-443.<br />

Castro-Gutierrez, N., McConnell, R., Andersson, K., Pacheco-Anton, F., and Hogstedt,<br />

C. (1997). Respiratory symptoms, spirometry and chronic occupational <strong>paraquat</strong><br />

exposure. Scand J Work Environ Health 23, 421-427.<br />

Castro, R., Prata, C., Oliveira, L., Carvalho, M. J., Santos, J., Carvalho, F., and<br />

Morgado, T. (2005). Paraquat <strong>into</strong>xication and hemocarboperfusion. Acta Med<br />

Port 18, 423-431.<br />

Çeçen, Ş. Ş., Cengiz, G., and Söylemezoğlu, T. (2002). The protective effects <strong>of</strong> Nacetylcysteine<br />

in <strong>paraquat</strong>-<strong>induced</strong> toxicity. J Fac Pharm Ankara 31, 259-271.<br />

Chan, B. S., Seale, J. P., and Duggin, G. C. (1997). The mechanism <strong>of</strong> excretion <strong>of</strong><br />

<strong>paraquat</strong> in rats. Toxicol Lett 15, 1-9.<br />

Chan, B. S. H., Lazzaro, V. A., Kirwin, P. D., Seale, J. P., and Duggin, G. G. (1996a).<br />

The effects <strong>of</strong> <strong>paraquat</strong> on two renal epi<strong>the</strong>lial cell lines-LLC-PK1 and MDCK.<br />

Res Comm Pharmacol Toxicol 1, 99-112.<br />

Chan, B. S. H., Lazzaro, V. A., Seale, J. P., and Duggin, G. C. (1996b).<br />

Characterization and uptake <strong>of</strong> <strong>paraquat</strong> by rat renal tubular cells in primary<br />

culture. Hum Exp Toxicol 15, 949-956.<br />

Chen, C. J., Chin, J. E., Ueda, K., Clark, D. P., Pastan, I., Gottesman, M. M., and<br />

Roninson, I. B. (1986 ). Internal duplication and homology with bacterial<br />

transport proteins in <strong>the</strong> mdr1 (P-glycoprotein) gene from multidrug-resistant<br />

human cells. Cell 47, 381-389.<br />

Chen, C. M., Fang, C. L., and Chang, C. H. (2001). Surfactant and corticosteroid effects<br />

on lung function in a rat model <strong>of</strong> acute lung injury. Crit Care Med 29, 2169-<br />

2175.<br />

Chen, C. M., Su, B., Hsu, C. C., and Wang, L. F. (2002a). Methylprednisolone does not<br />

enhance <strong>the</strong> surfactant effects on oxygenation and histology in <strong>paraquat</strong>-<strong>induced</strong><br />

rat lung injury. Intensive Care Med 28, 1138-1144.<br />

Chen, G. H., Lin, J. L., and Huang, Y. K. (2002b). Combined methylprednisolone and<br />

dexamethasone <strong>the</strong>rapy for <strong>paraquat</strong> poisoning. Crit Care Med 30, 2584-2587.<br />

Chen, K. W., Wu, M. H., Huang, J. J., and Yu, C. Y. (1994a). Bilateral spontaneous<br />

pneumothoraces, pneumopericardium, pneumomediastinum, and subcutaneous<br />

emphysema: a rare presentation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong>xication. Ann Emerg Med 23,<br />

1132-1134.<br />

Chen, N., Bowles, M. R., and Pond, S. M. (1992). Competition between <strong>paraquat</strong> and<br />

putrescine for uptake by suspensions <strong>of</strong> rat alveolar type II cells. Biochem<br />

Pharmacol 44, 1029-1036.<br />

Chen, N., Bowles, M. R., and Pond, S. M. (1994b). Prevention <strong>of</strong> <strong>paraquat</strong> toxicity in<br />

suspensions <strong>of</strong> alveolar type II cells by <strong>paraquat</strong>-specific antibodies. Hum Exp<br />

Toxicol 13, 551-557.<br />

Cheng, W., Fu, Y. X., Porres, J. M., Ross, D. A., and Lei, X. G. (1999). Seleniumdependent<br />

cellular glutathione peroxidase protects mice against a pro-oxidant-<br />

230


________________________________________________________Part IV – References<br />

<strong>induced</strong> oxidation <strong>of</strong> NADPH, NADH, lipids, and protein. FASEB J 13, 1467-<br />

1475.<br />

Cheng, W. H., Ho, Y. S., Valentine, B. A., Ross, D. A., Combs, G. F., Jr., and Lei, X.<br />

G. (1998). Cellular glutathione peroxidase is <strong>the</strong> mediator <strong>of</strong> body selenium to<br />

protect against <strong>paraquat</strong> lethality in transgenic mice. J Nutr 128, 1070-1076.<br />

Chester, G., and Ward, R. J. (1984). Occupational exposure and drift hazard during<br />

aerial application <strong>of</strong> <strong>paraquat</strong> to cotton. Arch Environ Contam Toxicol 13, 551-<br />

563.<br />

Cho, J. H., Yang, D. K., Kim, L., Ryu, J. S., Lee, H. L., Lim, C. M., and Koh, Y. S.<br />

(2005). Inhaled nitric oxide improves <strong>the</strong> survival <strong>of</strong> <strong>the</strong> <strong>paraquat</strong>-injured rats.<br />

Vascul Pharmacol 42, 171-178.<br />

Chollet, A., Muszynsky, J., Bismuth, C., Pham, J., Khouly, M., El , and Surugue, R.<br />

(1983). Hypo-oxygenation in <strong>paraquat</strong> poisoning. Apropos <strong>of</strong> 6 cases. Toxicol<br />

Eur Res 5, 71-75.<br />

Cingolani, C., Rogers, B., Lu, L., Kachi, S., Shen, J., and Campochiaro, P. A. (2006).<br />

Retinal degeneration from oxidative damage. Free Radic Biol Med 40, 660-669.<br />

Clark, D. G., McElligott, T. F., and Hurst, E. W. (1966). The toxicity <strong>of</strong> <strong>paraquat</strong>. Br J<br />

Ind Med 23, 126-132.<br />

Clejan, L., and Cederbaum, A. I. (1989). Synergistic interaction between NADPHcytochrome<br />

P-450 reductase, <strong>paraquat</strong> and iron in <strong>the</strong> generation <strong>of</strong> active<br />

oxygen radicals. Biochem Pharmacol 38, 1779-1786.<br />

Conning, D. M., Fletcher, K., and Swan, A. A. (1969). Paraquat and related bipyridyls.<br />

Br Med Bull 25, 245-249.<br />

Cordon-Cardo, C., O'Brien, J., Boccia, J., Casals, D., Bertino, J., and Melamed, M.<br />

(1990). Expression <strong>of</strong> <strong>the</strong> multidrug resistance gene product (P-glycoprotein) in<br />

human normal and tumor tissues. J Histochem Cytochem 38, 1277-1287.<br />

Crapo, J. D., Barry, B. E., Gehr, P., Bach<strong>of</strong>en, M., and Weibel, E. R. (1982). Cell<br />

number and cell characteristics <strong>of</strong> normal human lung. Am Rev Respir Dis 125,<br />

332-337.<br />

Daisley, H., Jr., and Simmons, V. (1999). Forensic analysis <strong>of</strong> acute fatal poisonings in<br />

<strong>the</strong> sou<strong>the</strong>rn districts <strong>of</strong> Trinidad. Vet Hum Toxicol 41, 23-25.<br />

Daniel, J. W., and Gage, J. C. (1966). Absorption and excretion <strong>of</strong> diquat and <strong>paraquat</strong><br />

in rats. Br J Ind Med 23, 133-136.<br />

Dargan, P. I., Wallace, C. I., and Jones, A. L. (2002). An evidence based flowchart to<br />

guide <strong>the</strong> management <strong>of</strong> acute salicylate (aspirin) overdose. Emerg Med J 19,<br />

206-209.<br />

Davies, D. S. (1987). Paraquat poisoning: The rationale for current treatment regimes.<br />

Hum Exp Toxicol 6, 37-40.<br />

Day, B. J., and Crapo, J. D. (1996). A metalloporhyrin superoxide dismutase mimetic<br />

protects against <strong>paraquat</strong>-<strong>induced</strong> lung injury in vivo. Toxicol Appl Pharmacol<br />

140, 94-100.<br />

Day, B. J., Patel, M., Calavetta, L., Chang, L. Y., and Stamler, J. S. (1999). A<br />

mechanism <strong>of</strong> <strong>paraquat</strong> toxicity involving nitric oxide synthase. Proc Natl Acad<br />

Sci U S A 96, 12760-12765.<br />

De Broe, M. E., Bismuth, C., De Groot, G., Heath, A., Okonek, S., Ritz, D. R.,<br />

Verpooten, G. A., Volans, G. N., and Widdop, B. (1986). Haemoperfusion: a<br />

useful <strong>the</strong>rapy for a severely poisoned patient? Hum Toxicol 5, 11-14.<br />

Dean, R. T., Fu, S., Stocker, R., and Davies, M. J. (1997). Biochemistry and pathology<br />

<strong>of</strong> radical-mediated protein oxidation. Biochem J 324, 1-18.<br />

231


Part IV – References____________________________________________________________________<br />

Dearden, L. C., Fairshter, R. D., Morrison, J. T., Wilson, A. F., and Brundage, M.<br />

(1982). Ultrastructural evidence <strong>of</strong> pulmonary capillary endo<strong>the</strong>lial damage<br />

from <strong>paraquat</strong>. Toxicology 24, 211-222.<br />

Demeule, M., Jodoin, J., Beaulieu, E., Brossard, M., and Beliveau, R. (1999).<br />

Dexamethasone modulation <strong>of</strong> multidrug transporters in normal tissues. FEBS<br />

Lett 442, 208-214.<br />

Deneke, S. M. (2000). Thiol-based antioxidants. Curr Top Cell Reg 36, 151–180.<br />

Devlin, C. M., Bowles, M. R., Gordon, R. B., and Pond, S. M. (1995). Production <strong>of</strong> a<br />

<strong>paraquat</strong>-specific murine single chain Fv fragment. J Biochem (Tokyo) 118, 480-<br />

487.<br />

Dey, M. S., Breeze, R. G., Hayton, W. L., Karara, A. H., and Krieger, R. I. (1990).<br />

Paraquat pharmacokinetics using a subcutaneous toxic low dose in <strong>the</strong> rat.<br />

Fundam. Appl. Toxicol 14, 208-216.<br />

Dicker, E., and Cederbaum, A. I. (1991). NADH-dependent generation <strong>of</strong> reactive<br />

oxygen species by microsomes in <strong>the</strong> presence <strong>of</strong> iron and redox cycling agents.<br />

Biochem Pharmacol 42, 529-535.<br />

Dinsdale, D., Preston, S. G., and Nemery, B. (1991). Effects <strong>of</strong> injury on [3H]putrescine<br />

uptake by types I and II cells in rat lung slices. Exp Mol Pathol 54, 218-229.<br />

Dodge, A. D. (1971). The mode <strong>of</strong> action <strong>of</strong> <strong>the</strong> bipyridylium herbicides, <strong>paraquat</strong> and<br />

diquat. Endeavour 30, 130-135.<br />

Dolan, M., Danzl, D. F., Horowitz, J., and Moore, D. (1984). Paraquat ingestion. Ann<br />

Emerg Med 13, 612-614.<br />

Douze, J. M., van Heyst, A. N., vanDijk, A., Maes, R. A., and Drost, R. H. (1975).<br />

Paraquat poisoning in man. Arch Toxicol 34, 129-136.<br />

Doyle, I. R., Hermans, C., Bernard, A., Nicholas, T. E., and Bersten, A. D. (1998).<br />

Clearance <strong>of</strong> Clara cell secretory protein (CC16) and surfactant proteins A and B<br />

from blood in acute respiratory failure. Am J Respir Crit Care Med 158, 1528-<br />

1535.<br />

Doyle, I. R., Nicholas, T. E., and Bersten, A. D. (1995). Serum surfactant protein-A<br />

levels in patients with acute cardiogenic pulmonary edema and adult respiratory<br />

distress syndrome. Am J Respir Crit Care Med 152, 307-317.<br />

Drault, J. N., Baelen, E., Mehdaoui, H., Delord, J. M., and Flament, F. (1999). Massive<br />

<strong>paraquat</strong> poisoning. Favorable course after treatment with n-acetylcysteine and<br />

early hemodialysis. Ann Fr Anesth Reanim 18, 534-537.<br />

Drew, R., Siddik, Z. H., and Gram, T. E. (1979). Uptake and efflux <strong>of</strong> 14C-<strong>paraquat</strong> by<br />

rat lung slices: <strong>the</strong> effect <strong>of</strong> imipramine and o<strong>the</strong>r drugs. Toxicol Appl<br />

Pharmacol 49, 473-478.<br />

Driesbach, R. H. (1983). In Handbook <strong>of</strong> Poisoning, pp. 118-122. Lange Medical<br />

Publications, Los Altos, CA.<br />

Dunbar, J. R., DeLucia, A. J., Acuff, R. V., and Ferslew, K. E. (1988). Prolonged,<br />

intravenous <strong>paraquat</strong> infusion in <strong>the</strong> rat. I. Failure <strong>of</strong> coinfused putrescine to<br />

attenuate pulmonary <strong>paraquat</strong> uptake, <strong>paraquat</strong>-<strong>induced</strong> biochemical changes, or<br />

lung injury. Toxicol Appl Pharmacol 94, 207-220.<br />

Dusinska, M., Kovacikova, Z., Vallova, B., and Collins, A. (1998). Responses <strong>of</strong><br />

alveolar macrophages and epi<strong>the</strong>lial type II cells to oxidative DNA damage<br />

caused by <strong>paraquat</strong>. Carcinogenesis 19, 809-812.<br />

Ecker, J. L., Hook, J. B., and Gibson, J. E. (1975). Nephrotoxicity <strong>of</strong> <strong>paraquat</strong> in mice.<br />

Toxicol Appl Pharmacol 34, 178-186.<br />

232


________________________________________________________Part IV – References<br />

Eddleston, M., Wilks, M. F., and Buckley, N. A. (2003). Prospects for treatment <strong>of</strong><br />

<strong>paraquat</strong>-<strong>induced</strong> lung fibrosis with immunosuppressive drugs and <strong>the</strong> need for<br />

better prediction <strong>of</strong> outcome: a systematic review. QJM 96, 809-824.<br />

Eggleston, L. V., and Krebs, H. A. (1974). Regulation <strong>of</strong> <strong>the</strong> pentose phosphate cycle.<br />

Biochem J 138, 425-435.<br />

Eisenman, A., Armali, Z., Raikhlin-Eisenkraft, B., Bentur, L., Bentur, Y., Guralnik, L.,<br />

and Enat, R. (1998). Nitric oxide inhalation for <strong>paraquat</strong>-<strong>induced</strong> lung injury. J<br />

Toxicol Clin Toxicol 36, 585-586.<br />

Evans, P., and Halliwell, B. (2001). Micronutrients: oxidant/antioxidant status. Br J<br />

Nutr 85, S67–S74.<br />

Fairshter, R. D., Dabir-Vaziri, N., Smith, W. R., Glauser, F. L., and Wilson, A. F.<br />

(1979). Paraquat poisoning: an analytical toxicologic study <strong>of</strong> three cases.<br />

Toxicology 12, 259-266.<br />

Farrington, J. A., Ebert, M., Land, E. J., and Fletcher, K. (1973). Bipyridylium<br />

quaternary salts and related compounds. V. Pulse radiolysis studies <strong>of</strong> <strong>the</strong><br />

reaction <strong>of</strong> <strong>paraquat</strong> radical with oxygen. Implications for <strong>the</strong> mode <strong>of</strong> action <strong>of</strong><br />

bipyridyl herbicides. Biochim Biophys Acta 314, 372-381.<br />

Farrington, J. A., Ledwith, A., and Stam, M. F. (1969). Cation-radicals: Oxidation <strong>of</strong><br />

methoxide ion with 1,1' -dimethyl-4,4' -bipyridylium dichloride (<strong>paraquat</strong><br />

dichloride). J Chem Soc Chem Commun, 259-260.<br />

Fell, A. F., Jarvie, D. R., and Stewart, M. J. (1981). Analysis for <strong>paraquat</strong> by second-<br />

and fourth-derivative spectroscopy. Clin Chem 27, 286-292.<br />

Fennelly, J. J., Fitzgerald, M. X., and Fitzgerald, O. (1971). Recovery from severe<br />

<strong>paraquat</strong> poisoning following forced diuresis and immunosuppressive <strong>the</strong>rapy. J<br />

Ir Med Assoc 64, 69-71.<br />

Fennelly, J. J., Gallagher, J. T., and Carroll, R. T. (1968). Paraquat poisoning in a<br />

pregnant woman. Br Med J 3, 722-723.<br />

Ferguson, D. M. (1971). Renal handling <strong>of</strong> <strong>paraquat</strong>. Br J Pharmacol 42, 636P.<br />

Fisher, H. K., Clements, J. A., Tierney, D. F., and Wright, R. R. (1975). Pulmonary<br />

effects <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> first day after injection. Am J Physiol 228, 1217-1223.<br />

Fisher, H. K., Clements, J. A., and Wright, R. R. (1973). Enhancement <strong>of</strong> oxygen<br />

toxicity by <strong>the</strong> herbicide <strong>paraquat</strong>. Am Rev Respir Dis 107, 246-252.<br />

Florkowski, C. M., Bradberry, S. M., Ching, G. W., and Jones, A. F. (1992). Acute<br />

renal failure in a case <strong>of</strong> <strong>paraquat</strong> poisoning with relative absence <strong>of</strong> pulmonary<br />

toxicity. Postgrad Med J 68, 660-662.<br />

Fox, D. A., and McCune, W. J. (1994). Immunosuppressive drug <strong>the</strong>rapy <strong>of</strong> systemic<br />

lupus ery<strong>the</strong>matosus. Rheum Dis Clin North Am 20, 265-299.<br />

Frank, L., Neriishi, K., Sio, R., and Pascual, D. (1982). Protection from <strong>paraquat</strong><strong>induced</strong><br />

lung damage and lethality in adult rats pretreated with cl<strong>of</strong>ibrate.<br />

Toxicol Appl Pharmacol 66, 269-277.<br />

Freeman, B. A., Turrens, J. F., Mirza, Z., Crapo, J. D., and Young, S. L. (1985).<br />

Modulation <strong>of</strong> oxidant lung injury by using liposome-entrapped superoxide<br />

dismutase and catalase. Fed Proc 44, 2591-2595.<br />

Fritz, K. L., Nelson, T. L., Ruiz-Velasco, V., and Mercurio, S. D. (1994). Acute<br />

intramuscular injection <strong>of</strong> oils or <strong>the</strong> oleic acid component protects mice against<br />

<strong>paraquat</strong> lethality. J Nutr 124, 425-429.<br />

Fuke, C., Ameno, K., Ameno, S., Kiriu, T., Shinohara, T., Sogo, K., and Ijiri, I. (1992).<br />

A rapid, simultaneous determination <strong>of</strong> <strong>paraquat</strong> and diquat in serum and urine<br />

using second-derivative spectroscopy. J Anal Toxicol 16, 214-216.<br />

233


Part IV – References____________________________________________________________________<br />

Fukuda, Y., Ferrans, V. J., Schoenberger, C. I., Rennard, S. I., and Crystal, R. G.<br />

(1985). Patterns <strong>of</strong> pulmonary structural remodeling after experimental <strong>paraquat</strong><br />

toxicity. The morphogenesis <strong>of</strong> intraalveolar fibrosis. Am J Pathol 118, 452-475.<br />

Fukushima, T., Yamada, K., Isobe, A., Shiwaku, K., and Yamane, Y. (1993).<br />

Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. I. NADH oxidation and <strong>paraquat</strong> radical<br />

formation via complex I. Exp Toxicol Pathol 45, 345-349.<br />

Gadek, J. E., Fells, G. A., Zimmerman, R. L., and Crystal, R. G. (1984). Role <strong>of</strong><br />

connective tissue proteases in <strong>the</strong> pathogenesis <strong>of</strong> chronic inflammatory lung<br />

disease. Environ Health Perspect 55, 297-306.<br />

Gardiner, A. J. (1972). Pulmonary oedema in paraguat poisoning. Thorax, 132-135.<br />

Garnier, R., Chataigner, D., Efthymiou, M. L., Moraillon, I., and Bramary, F. (1994).<br />

Paraquat poisoning by skin absorption: report <strong>of</strong> two cases. Vet Hum Toxicol 36,<br />

313-315.<br />

Gaudreault, P., Friedman, P. A., and Lovejoy, F. H., Jr. (1985). Efficacy <strong>of</strong> activated<br />

charcoal and magnesium citrate in <strong>the</strong> treatment <strong>of</strong> oral <strong>paraquat</strong> <strong>into</strong>xication.<br />

Ann Emerg Med 14, 123-125.<br />

Gharaee-Kermani, M., and Phan, S. H. (2005). Molecular <strong>mechanisms</strong> <strong>of</strong> and possible<br />

treatment strategies for idiopathic pulmonary fibrosis. Curr Pharm Des 11,<br />

3943-3971.<br />

Ghazi-Khansari, M., Mohammadi-Karakani, A., Sotoudeh, M., Mokhtary, P., Pour-<br />

Esmaeil, E., and Maghsoud, S. (2007). Antifibrotic effect <strong>of</strong> captopril and<br />

enalapril on <strong>paraquat</strong>-<strong>induced</strong> lung fibrosis in rats. J Appl Toxicol, DOI:<br />

10.1002/jat.1212.<br />

Ghazi-Khansari, M., Nasiri, G., and Honarjoo, M. (2005). Decreasing <strong>the</strong> oxidant stress<br />

from <strong>paraquat</strong> in isolated perfused rat lung using captopril and niacin. Arch<br />

Toxicol 79, 341-345.<br />

Gianetti, J., Bevilacqua, S., and Caterina, R. D. (2002). Inhaled nitric oxide: more than a<br />

selective pulmonary vasodilator. Eur J Clin Invest 32, 628-635.<br />

Glass, M., Su<strong>the</strong>rland, M. W., Forman, H. J., and Fisher, A. B. (1985). Selenium<br />

deficiency potentiates <strong>paraquat</strong>-<strong>induced</strong> lipid peroxidation in isolated perfused<br />

rat lung. J Appl Physiol 59, 619-622.<br />

Goldenberg, H., Huttinger, M., Kampfer, P., Kramar, R., and Pavelka, M. (1976). Effect<br />

<strong>of</strong> cl<strong>of</strong>ibrate application on morphology and enzyme content <strong>of</strong> liver<br />

peroxisomes. Histochemistry 46, 189-196.<br />

Gordonsmith, R. H., Brooke-Taylor, S., Smith, L. L., and Cohen, G. M. (1983).<br />

Structural requirements <strong>of</strong> compounds to inhibit pulmonary diamine<br />

accumulation. Biochem Pharmacol 32, 3701-3709.<br />

Gottesman, M. M., and Pastan, I. (1993). Biochemistry <strong>of</strong> multidrug resistance<br />

mediated by <strong>the</strong> multidrug transporter. Annu Rev Biochem 62, 385-427.<br />

Groce, D. F., and Kimbrough, R. D. (1982). Acute and subacute toxicity in Sherman<br />

strain rats exposed to 4,4'- and 2,2'-dipyridyl. J Toxicol Environ Health A 10,<br />

363-372.<br />

Grover, R., Smith, A. E., and Korven, H. C. (1980). A comparison <strong>of</strong> chemical and<br />

cultural control <strong>of</strong> weeds in irrigation ditchbanks. Can J Plant Sci 60, 185-195.<br />

Hagen, T. M., Brown, L. A., and Jones, D. P. (1986). Protection against <strong>paraquat</strong><strong>induced</strong><br />

injury by exogenous GSH in pulmonary alveolar type II cells. Biochem<br />

Pharmacol 35, 4537-4542.<br />

Haley, T. J. (1979). Review <strong>of</strong> <strong>the</strong> toxicology <strong>of</strong> <strong>paraquat</strong> (1,1-dimethyl-4,4dipyridylium<br />

chloride). Clin Toxicol 14, 1-46.<br />

234


________________________________________________________Part IV – References<br />

Halliwell, B. (1996). Vitamin C: antioxidant or pro-oxidant in vivo? Free Rad Res 25,<br />

439-454.<br />

Hampson, E. C., and Pond, S. M. (1988). Failure <strong>of</strong> haemoperfusion and haemodialysis<br />

to prevent death in <strong>paraquat</strong> poisoning. A retrospective review <strong>of</strong> 42 patients.<br />

Med Toxicol Adverse Drug Exp 3, 64-71.<br />

Hance, R. L., Byast, T. H., and Smith, P. D. (1980). Apparent decomposition <strong>of</strong><br />

<strong>paraquat</strong> in soil. Soil Biol Biochem 12, 447-448.<br />

Hantson, P., Weynand, B., Doyle, I., Bernand, A., and Hermans, C. (2006).<br />

Pneumoproteins as markers <strong>of</strong> <strong>paraquat</strong> lung injury: A clinical case. J Clin<br />

Forensic Med, doi:10.1016/j.jcfm.2006.1009.1003.<br />

Harley, J. B., Grinspan, S., and Root, R. K. (1977). Paraquat suicide in a young woman:<br />

Results <strong>of</strong> <strong>the</strong>rapy directed against <strong>the</strong> superoxide radical. Yale J Biol Med 50,<br />

481-488.<br />

Hart, C. M. (1999). Nitric Oxide in Adult Lung Disease. Chest 115, 1407 - 1417.<br />

Hart, T. B. (1987). Paraquat--a review <strong>of</strong> safety in agricultural and horticultural use.<br />

Hum Toxicol 6, 13-18.<br />

Hart, T. B., Nevitt, A., and Whitehead, A. (1984). A new statistical approach to <strong>the</strong><br />

prognostic significance <strong>of</strong> plasma concentrations. Lancet 2, 1222-1223.<br />

Hawksworth, G. M., Bennett, P. N., and Davies, D. S. (1981). Kinetics <strong>of</strong> <strong>paraquat</strong><br />

elimination in <strong>the</strong> dog. Toxicol Appl Pharmacol 57, 139-145.<br />

Hearn, C. E., and Keir, W. (1971). Nail damage in spray operators exposed to <strong>paraquat</strong>.<br />

Br J Ind Med 28, 399-403.<br />

Heby, O. (1981). Role <strong>of</strong> polyamines in <strong>the</strong> control <strong>of</strong> cell proliferation and<br />

differentiation. Differentiation 19, 1-20.<br />

Hermans, C., and Bernard, A. (1998). Pneumoproteinemia: a new perspective in <strong>the</strong><br />

assessment <strong>of</strong> lung disorders. Eur Respir J 11, 801-803.<br />

Hermans, C., and Bernard, A. (1999). Lung epi<strong>the</strong>lium specific proteins: characteristics<br />

and potential applications as markers. State <strong>of</strong> <strong>the</strong> art. Am J Respir Crit Care<br />

Med 159, 646-678.<br />

Hoet, P. H., Dinsdale, M. D., Lewis, C. P. L., Verbeken, E. K., Lauweryns, J. M., and<br />

Nemery, B. (1993). Kinetics and cellular localisation <strong>of</strong> putrescine uptake in<br />

human lung. Thorax 48, 1235-1241.<br />

Hoet, P. H., Lewis, C. P., Demedts, M., and Nemery, B. (1994). Putrescine and <strong>paraquat</strong><br />

uptake in human lung slices and isolated human type II pneumocytes. Biochem.<br />

Pharmacol 48, 517-524.<br />

Hoet, P. H., Lewis, C. P., Dinsdale, D., Demedts, M., and Nemery, B. (1995).<br />

Putrescine uptake in hamster lung slices and primary cultures <strong>of</strong> type II<br />

pneumocytes. Am J Physiol 269, L681-L689.<br />

H<strong>of</strong>fer, E., Avidor, I., Benjaminov, O., Shenker, L., Tabak, A., Tamir, A., Merzbach,<br />

D., and Taitelman, U. (1993). N-acetylcysteine delays <strong>the</strong> infiltration <strong>of</strong><br />

inflammatory cells <strong>into</strong> <strong>the</strong> lungs <strong>of</strong> <strong>paraquat</strong>-<strong>into</strong>xicated rats. Toxicol Appl<br />

Pharmacol 120, 8-12.<br />

H<strong>of</strong>fer, E., Baum, Y., Tabak, A., and Taitelman, U. (1996). N-Acetylcysteine increases<br />

<strong>the</strong> glutathione content and protects rat alveolar type II cells against <strong>paraquat</strong><strong>induced</strong><br />

cytotoxicity. Toxicol Lett 84, 7-12.<br />

H<strong>of</strong>fer, E., and Taitelman, U. (1989). Exposure to <strong>paraquat</strong> through skin absorption:<br />

clinical and laboratory observations <strong>of</strong> accidental splashing on healthy skin <strong>of</strong><br />

agricultural workers. Hum Toxicol 8, 483-485.<br />

Homer, R. F., Mees, G. C., and Tomlinson, T. E. (1960). Mode <strong>of</strong> action <strong>of</strong> dipyridyl<br />

quaternary salts as herbicides. J Sci Food Agric 11, 309-315.<br />

235


Part IV – References____________________________________________________________________<br />

Homer, R. F., and Tomlinson, T. E. (1959). Redox Properties <strong>of</strong> some Dipyridyl<br />

Quaternary Salts. Nature 184, 2012 - 2013.<br />

Hong, S. F., Yang, J. O., Lee, E. Y., and Kim, S. H. (2003). Effect <strong>of</strong> haemoperfusion<br />

on plasma <strong>paraquat</strong> concentration in vitro and in vivo. Toxicol Ind Health, 17-<br />

23.<br />

Hong, S. Y., Gil, H. W., Yang, J. O., Lee, E. Y., Na, J. O., Seo, K. H., and Kim, Y. H.<br />

(2005). Clinical implications <strong>of</strong> <strong>the</strong> ethane in exhaled breath in patients with<br />

acute <strong>paraquat</strong> <strong>into</strong>xication. Chest 128, 1506-1510.<br />

Hong, S. Y., Hwang, K. Y., Lee, E. Y., Eun, S. W., Cho, S. R., Han, C. S., Park, Y. H.,<br />

and Chang, S. K. (2002). Effect <strong>of</strong> vitamin C on plasma total antioxidant status<br />

in patients with <strong>paraquat</strong> <strong>into</strong>xication. Toxicol Lett 126, 51-59.<br />

Houze, P., Baud, F. J., Mouy, R., Bismuth, C., Bourdon, R., and Scherrmann, J. M.<br />

(1990). Toxicokinetics <strong>of</strong> <strong>paraquat</strong> in humans. Hum Exp Toxicol 9, 5-12.<br />

Houze, P., Baud, F. J., and Scherrmann, J. M. (1995). Toxicokinetics <strong>of</strong> <strong>paraquat</strong>. In<br />

Paraquat poisoning: <strong>mechanisms</strong>, prevention, treatment (C. Bismuth, and A. H.<br />

Hall, Eds.), pp. 161-193. Marcel Dekker, New York.<br />

Howard, J. K. (1983). The myth <strong>of</strong> <strong>paraquat</strong> inhalation as a route for human poisoning.<br />

J Toxicol Clin Toxicol 20, 191-193.<br />

Howard, J. K., Sabapathy, N. N., and Whitehead, P. A. (1981). A study <strong>of</strong> <strong>the</strong> health <strong>of</strong><br />

Malaysian plantation workers with particular reference to <strong>paraquat</strong> spraymen. Br<br />

J Ind Med 38, 110-116.<br />

Hsu, H. H., Chang, C. T., and Lin, J. L. (2003). Intravenous <strong>paraquat</strong> poisoning-<strong>induced</strong><br />

<strong>multiple</strong> <strong>organ</strong> failure and fatality--a report <strong>of</strong> two cases. J Toxicol Clin Toxicol<br />

41, 87-90.<br />

Huang, N. C., Hung, Y. M., Lin, S. L., Wann, S. R., Hsu, C. W., Ger, L. P., Hung, S.<br />

Y., Chung, H. M., and Yeh, J. H. (2006). Fur<strong>the</strong>r evidence <strong>of</strong> <strong>the</strong> usefulness <strong>of</strong><br />

Acute Physiology and Chronic Health Evaluation II scoring system in acute<br />

<strong>paraquat</strong> poisoning. Clin Toxicol (Phila) 44, 99-102.<br />

Huang, N. C., Lin, S. L., Hung, Y. M., Hung, S. Y., and Chung, H. M. (2003). Severity<br />

assessment in acute <strong>paraquat</strong> poisoning by analysis <strong>of</strong> APACHE II score. J<br />

Formos Med Assoc 102, 782-787.<br />

Hudson, M., Patel, S. B., Ewen, S. W., Smith, C. C., and Friend, J. A. (1991). Paraquat<br />

<strong>induced</strong> pulmonary fibrosis in three survivors. Thorax 46, 201-204.<br />

Hughes, R. D., Millburn, P., and Williams, R. T. (1973). Biliary excretion <strong>of</strong> some<br />

diquaternary ammonium cations in <strong>the</strong> rat, guinea pig and rabbit. Biochem J 136,<br />

979-984<br />

Huh, J. W., Hong, S. B., Lim, C. M., Do, K. H., Lee, J. S., and Koh, Y. (2006).<br />

Sequential radiologic and functional pulmonary changes in patients with<br />

<strong>paraquat</strong> <strong>into</strong>xication. Int J Occup Environ Health 12, :203-208.<br />

Ikebuchi, J., Proudfoot, A. T., Matsubara, K., Hampson, E. C., Tomita, M., Suzuki, K.,<br />

Fuke, C., Ijiri, I., Tsunerari, T., Yuasa, I., and Okada, K. (1993). Toxicological<br />

index <strong>of</strong> <strong>paraquat</strong>: a new strategy for assessment <strong>of</strong> severity <strong>of</strong> <strong>paraquat</strong><br />

poisoning in 128 patients. Forensic Sci Int 59, 85-87.<br />

Ilett, K. F., Stripp, B., Menard, R. H., Reid, W. D., and Gillette, J. R. (1974). Studies on<br />

<strong>the</strong> mechanism <strong>of</strong> <strong>the</strong> lung toxicity <strong>of</strong> <strong>paraquat</strong>: comparison <strong>of</strong> tissue distribution<br />

and some biochemical parameters in rats and rabbits. Toxicol Appl Pharmacol<br />

28, 216-226.<br />

Im, J. G., Lee, K. S., Han, M. C., Kim, S. J., and Kim, I. O. (1991). Paraquat poisoning:<br />

findings on chest radiography and CT in 42 patients. AJR Am J Roentgenol 157,<br />

697-701.<br />

236


________________________________________________________Part IV – References<br />

Ishii, K., Adachi, J., Tomita, M., Kurosaka, M., and Ueno, Y. (2002). Oxysterols as<br />

indices <strong>of</strong> oxidative stress in man after <strong>paraquat</strong> ingestion. Free Radic Res 36,<br />

163-168.<br />

Ito, M., and Kuwana, T. (1971). J Electroanal Chem Interfacial Electrochem 32, 415.<br />

Jaiswal, R., Khan, M. A., and Musarrat, J. (2002). Photosensitized <strong>paraquat</strong>-<strong>induced</strong><br />

structural alterations and free radical mediated fragmentation <strong>of</strong> serum albumin.<br />

J Photochem Photobiol B 67, 163-170.<br />

Janne, J., Poso, H., and Raina, A. (1978). Polyamines in rapid growth and cancer.<br />

Biochim Biophys Acta 473, 241-293.<br />

Jarvie, D. R., Fell, A. F., and Stewart, M. J. (1981). A rapid method for <strong>the</strong> emergency<br />

analysis <strong>of</strong> <strong>paraquat</strong> in plasma using a second derivative spectroscopy. Clin<br />

Chim Acta 117, 153-165.<br />

Jenq, C. C., Wu, C. D., and Lin, J. L. (2005). Mo<strong>the</strong>r and fetus both survive from severe<br />

<strong>paraquat</strong> <strong>into</strong>xication. Clin Toxicol (Phila) 43, 291-295.<br />

Johnston, S. C., Bowles, M., Winzor, D. J., and Pond, S. M. (1988). Comparison <strong>of</strong><br />

<strong>paraquat</strong>-specific murine monoclonal antibodies produced by in vitro and in<br />

vivo immunization. Fundam Appl Toxicol 11, 261-267.<br />

Jones, A. L., Elton, R., and Flanagan, R. (1999). Multiple logistic regression analysis <strong>of</strong><br />

plasma <strong>paraquat</strong> concentrations as a predictor <strong>of</strong> outcome in 375 cases <strong>of</strong><br />

<strong>paraquat</strong> poisoning. QJM 92, 573-578.<br />

Jones, G. M., and Vale, J. A. (2000). Mechanisms <strong>of</strong> toxicity, clinical features, and<br />

management <strong>of</strong> diquat poisoning: a review. Clin Toxicol 38, 123-128.<br />

Kang, S. A., Jang, Y. J., and Park, H. (1998). In vivo dual effects <strong>of</strong> vitamin C on<br />

<strong>paraquat</strong>-<strong>induced</strong> lung damage: dependence on released metals from <strong>the</strong><br />

damaged tissue. Free Rad Res 28, 93-107.<br />

Kao, C. H., Hsieh, J. F., Ho, Y. J., Hung, D. Z., Lin, T. J., and Ding, H. J. (1999). Acute<br />

<strong>paraquat</strong> <strong>into</strong>xication: using nuclear pulmonary studies to predict patient<br />

outcome. Chest 116, 709-714.<br />

Kaojarern, S., and Ongphiphadhanakul, B. (1991). Predicting outcomes in <strong>paraquat</strong><br />

poisonings. Vet Hum Toxicol 33, 115-118.<br />

Karl, P. I., and Friedman, P. A. (1983). Competition between <strong>paraquat</strong> and putrescine<br />

for accumulation by rat lung slices. Toxicology 26, 317-323.<br />

Kazui, M., Andreoni, K. A., Norris, E. J., Klein, A. S., Burdick, J. F., Beattie, C.,<br />

Sehnert, S. S., Bell, W. R., Bulkley, G. B., and Risby, T. H. (1992). Breath<br />

ethane: a specific indicator <strong>of</strong> free-radical-mediated lipid peroxidation following<br />

reperfusion <strong>of</strong> <strong>the</strong> ischemic liver. Free Radic Biol Med 13, 509-515.<br />

Keeling, P. L., and Smith, L. L. (1982). Relevance <strong>of</strong> NADPH depletion and mixed<br />

disulphide formation in rat lung to <strong>the</strong> mechanism <strong>of</strong> cell damage following<br />

<strong>paraquat</strong> administration. Biochem Pharmacol 31, 3243-3249.<br />

Keeling, P. L., Smith, L. L., and Aldridge, W. N. (1982). The formation <strong>of</strong> mixed<br />

disulphides in rat lung following <strong>paraquat</strong> administration. Correlation with<br />

changes in intermediary metabolism. Biochim Biophys Acta 716, 249-257.<br />

Kehrer, J. P., Haschek, W. M., and Witschi, H. (1979). The influence <strong>of</strong> hyperoxia on<br />

<strong>the</strong> acute toxicity <strong>of</strong> <strong>paraquat</strong> and diquat. Drug Chem Toxicol 2, 397-408.<br />

Kelly, G. S. (1998). Clinical applications <strong>of</strong> N-acetylcysteine. Altern Med Rev 3, 114-<br />

127.<br />

Kelner, M. J., Bagnell, R., Hale, B., and Alexander, N. M. (1988). Methylene blue<br />

competes with <strong>paraquat</strong> for reduction by flavo-enzymes resulting in decreased<br />

superoxide production in <strong>the</strong> presence <strong>of</strong> heme proteins. Arch Biochem Biophys<br />

262, 422-426.<br />

237


Part IV – References____________________________________________________________________<br />

Kerr, F., Patel, A. R., Scott, P. D., and Tompsett, S. L. (1968). Paraquat poisoning<br />

treated by forced diuresis. Br Med J 3, 290-291.<br />

Kimbrough, R. D., and Gaines, T. B. (1970). Toxicity <strong>of</strong> <strong>paraquat</strong> to rats and its effect<br />

on rat lungs. Toxicol Appl Pharm 17, 679-690.<br />

Kitazawa, Y., Matsubara, M., Takeyama, N., and Tanaka, T. (1991). The role <strong>of</strong><br />

xanthine oxidase in <strong>paraquat</strong> <strong>into</strong>xication. Arch Biochem Biophys 288, 220-224.<br />

Knaus, W. A., Draper, E. A., Wagner, D. P., and Zimmerman, J. E. (1985). APACHE<br />

II: a severity <strong>of</strong> disease classification system. Crit Care Med 13, 818-829.<br />

Kneepkens, C. M., Lepage, G., and Roy, C. C. (1994). The potential <strong>of</strong> <strong>the</strong> hydrocarbon<br />

breath test as a measure <strong>of</strong> lipid peroxidation. Free Radic Biol Med 17, 127-160.<br />

Kohen, R., and Chevion, M. (1985a). Paraquat toxicity is enhanced by iron and reduced<br />

by desferrioxamine in laboratory mice. Biochem Pharmacol 34, 1841-1843.<br />

Kohen, R., and Chevion, M. (1985b). Paraquat toxicity is mediated by transition metals.<br />

In Biological and In<strong>organ</strong>ic Copper Chemistry (K. D. Karlin, and J. Zubieta,<br />

Eds.), pp. 159-172. Adenine Press, New York.<br />

Kohen, R., and Chevion, M. (1985c). Transition metals potentiate <strong>paraquat</strong> toxicity.<br />

Free Radic Res Commun 1, 79-88.<br />

Kojima, S., Miyazaki, Y., Honda, T., Kiyozumi, M., Shimada, H., and Funakoshi, T.<br />

(1992). Effect <strong>of</strong> L-cystine on toxicity <strong>of</strong> <strong>paraquat</strong> in mice. Toxicol Lett 60, 75-<br />

82.<br />

Koppel, C., von Wissmann, C., Barckow, D., Rossaint, R., Falke, K., Stoltenburg-<br />

Didinger, G., and Schnoy, N. (1994). Inhaled nitric oxide in advanced <strong>paraquat</strong><br />

<strong>into</strong>xication. J Toxicol Clin Toxicol 32, 205-214.<br />

Kosower, E. M., and Cotter, J. L. (1964). The reduction <strong>of</strong> 1-methyl-4-cyanopyridinium<br />

ion to methyl viologen cation radical. J Am Chem Soc 86, 5524-5527.<br />

Koyama, K., Yamashita, M., Tai, T., Tajima, K., Mizutani, T., and Naito, H. (1987).<br />

The effect <strong>of</strong> chlorpromazine on <strong>paraquat</strong> poisoning in rats. Vet Hum Toxicol 29,<br />

117-121.<br />

Kumagai, J., and Johnson, L. R. (1988). Characteristics <strong>of</strong> putrescine uptake in isolated<br />

rat enterocytes. Am J Physiol 254, G81-G86.<br />

Kuo, T. L., Lin, D. L., Liu, R. H., Moriya, F., and Hashimoto, Y. (2001). Spectra<br />

interference between diquat and <strong>paraquat</strong> by second derivative<br />

spectrophotometry. Forensic Sci Int 121, 134-139.<br />

Kurisaki, E. (1985). Lipid peroxidation in human <strong>paraquat</strong> poisoning. J Toxicol Sci 10,<br />

29-33.<br />

Kuroki, Y., Takahashi, H., Chiba, H., and Akino, T. (1998). Surfactant proteins A and<br />

D: disease markers. Biochim Biophys Acta 1408, 334-345.<br />

Landrigan, P. J., Powell, K. E., James, L. M., and Taylor, P. R. (1983). Paraquat and<br />

marijuana: epidemiologic risk assessment. Am J Public Health 73, 784-788.<br />

Lang, J. D., McArdle, P. J., O'Reilly, P. J., and Matalon, S. (2002). Oxidant-antioxidant<br />

balance in acute lung injury. Chest 122, 314S-320S.<br />

Lasky, L. A., and Ortiz, L. A. (2001). Antifibrotic <strong>the</strong>rapy for <strong>the</strong> treatment <strong>of</strong><br />

pulmonary fibrosis. Am J Med Sci 332, 213-218.<br />

Lazo, J. S., Kondo, Y., Dellapiazza, D., Michalska, A. E., Choo, K. H., and Pitt, B. R.<br />

(1995). Enhanced sensitivity to oxidative stress in cultured embryonic cells from<br />

transgenic mice deficient in metallothionein I and II genes. J Biol Chem 270,<br />

5506-5510.<br />

Ledwith, A., and Woods, H. J. (1970). Charge transfer spectra and reaction<br />

intermediates. Part II. Stable crystalline complexes from phenols and NNdimethyl-4,4-bipyridylium<br />

(<strong>paraquat</strong>) salts. J Chem Soc C, 1422 - 1425.<br />

238


________________________________________________________Part IV – References<br />

Lee, E. Y., Hwang, K. Y., Yang, J. O., and Hong, S. Y. (2002). Predictors <strong>of</strong> survival<br />

after acute <strong>paraquat</strong> poisoning. Toxicol Ind Health 18, 201-206.<br />

Lee, S. H., Lee, K. S., Ahn, J. M., Kim, S. H., and Hong, S. Y. (1995). Paraquat<br />

poisoning <strong>of</strong> <strong>the</strong> lung: thin-section CT findings. Radiology 195, 271-274.<br />

Lewis, C. P., Haschek, W. M., Wyatt, I., Cohen, G. M., and Smith, L. L. (1989). The<br />

accumulation <strong>of</strong> cystamine and its metabolism to taurine in rat lung slices.<br />

Biochem Pharmacol 38, 481-488.<br />

Lheureux, P., Leduc, D., Vanbinst, R., and Askenasi, R. (1995). Survival in a case <strong>of</strong><br />

massive <strong>paraquat</strong> ingestion. Chest 107, 285-289.<br />

Lin, J. L., Leu, M. L., Liu, Y. C., and Chen, G. H. (1999). A prospective clinical trial <strong>of</strong><br />

pulse <strong>the</strong>rapy with glucocorticoid and cyclophosphamide in moderate to severe<br />

<strong>paraquat</strong>-poisoned patients. Am J Respir Crit Care Med 159, 357-360.<br />

Lin, J. L., Lin-Tan, D. T., Chen, K. H., and Huang, W. H. (2006). Repeated pulse <strong>of</strong><br />

methylprednisolone and cyclophosphamide with continuous dexamethasone<br />

<strong>the</strong>rapy for patients with severe <strong>paraquat</strong> poisoning. Crit Care Med 34, 368-373.<br />

Lin, J. L., Liu, L., and Leu, M. L. (1995). Recovery <strong>of</strong> respiratory function in survivors<br />

with <strong>paraquat</strong> <strong>into</strong>xication. Arch Environ Health 50, 432-439.<br />

Lin, J. L., Wei, M. C., and Liu, Y. C. (1996). Pulse <strong>the</strong>rapy with cyclophosphamide and<br />

methylprednisolone in patients with moderate to severe <strong>paraquat</strong> poisoning: a<br />

preliminary report. Thorax 51, 661-663.<br />

Lin, N. C., Lin, J. L., Lin-Tan, D. T., and Yu, C. C. (2003). Combined initial<br />

cyclophosphamide with repeated methylprednisolone pulse <strong>the</strong>rapy for severe<br />

<strong>paraquat</strong> poisoning from dermal exposure. J Toxicol Clin Toxicol 41, 877-881.<br />

Ling, V., Kartner, N., Sudo, T., Siminovitch, L., and Riordan, J. R. (1983). Multidrugresistance<br />

phenotype in Chinese hamster ovary cells. Cancer Treat Rep 67, 869-<br />

874.<br />

Liochev, S. I., and Fridovich, I. (1994). Paraquat diaphorases in Escherichia coli. Free<br />

Radical Biol Med 16, 555-559.<br />

Lock, E. A., and Ishmael, J. (1979). The acute toxic effects <strong>of</strong> <strong>paraquat</strong> and diquat on<br />

<strong>the</strong> rat kidney. Toxicol Appl Pharmacol 50, 67-76.<br />

Lock, E. A., Smith, L. L., and Rose, M. S. (1976). Inhibition <strong>of</strong> <strong>paraquat</strong> accumulation<br />

in rat lung slices by a component <strong>of</strong> rat plasma and a variety <strong>of</strong> drugs and<br />

endogenous amines. Biochem Pharmacol 25, 1769-1772.<br />

Longstaffe, J. A., Humphreys, D. J., Hayward, A. H. S., and Stodulski, J. B. J. (1981).<br />

Paraquat poisoning in dogs and cats-dffferences between accidental and<br />

malicious poisoning. J Small Anim. Pract 22, 153-156.<br />

Lopez Lago, A. M., Rivero Velasco, C., Galban Rodriguez, C., Marino Rozados, A.,<br />

Pineiro Sande, N., and Ferrer Vizoso, E. (2002). Paraquat poisoning and<br />

hemoperfusion with activated charcoal. An Med Interna 19, 310-312.<br />

Lugo-Vallin Ndel, V., Pascuzzo-Lima, C., Maradei-Irastorza, I., Ramirez-Sanchez, M.,<br />

Sosa-Sequera, M., Aguero-Pena, R., and Granado-Duque, A. (2002). Prolonged<br />

survival after experimental <strong>paraquat</strong> <strong>into</strong>xication: role <strong>of</strong> alternative<br />

antioxidants. Vet Hum Toxicol 44, 40-41.<br />

Malmgvist, E., Grossman, G., Ivemark, B., and Robertson, B. (1973). Pulmonary<br />

phospholipids and surface properties <strong>of</strong> alveolar wall in experimental <strong>paraquat</strong><br />

poisoning. Scand J Resp Dis 54, 206-214.<br />

Malone, J. D., Carmody, M., Keogh, B., and O'Dwyer, W. F. (1971). Paraquat<br />

poisoning; a review <strong>of</strong> nineteen cases. J Ir Med Assoc 64, 59-68.<br />

Manabe, J., and Ogata, T. (1987). Lung fibrosis <strong>induced</strong> by diquat after intratracheal<br />

administration. Arch Toxicol 60, 427-431.<br />

239


Part IV – References____________________________________________________________________<br />

Marczynski, B. T. (1985). The binding <strong>of</strong> spermine to polynucleotides and<br />

complementary oligonucleotides at near physiological ionic strength. Int. J.<br />

Biochem 17.<br />

Maruyama, K., Takeuchi, M., Chikusa, H., and Muneyuki, M. (1995). Reduction <strong>of</strong><br />

intrapulmonary shunt by low-dose inhaled nitric oxide in a patient with latestage<br />

respiratory distress associated with <strong>paraquat</strong> poisoning. Intensive Care<br />

Med 21, 778-779.<br />

Masek, L., and Richards, R. J. (1990). Interactions between <strong>paraquat</strong>, endogenous lung<br />

amines' antioxidants and isolated mouse Clara cells. Toxicology 63, 315-326.<br />

Matsubara, M., Yamagami, K., Kitazawa, Y., Kawamoto, K., and Tanaka, T. (1996).<br />

Paraquat causes S-phase arrest <strong>of</strong> rat liver and lung cells in vivo. Arch Toxicol<br />

70, 514-518.<br />

Mat<strong>the</strong>w, H., Logan, A., Woodruff, M. F., and Heard, B. (1968). Paraquat poisoning-lung<br />

transplantation. Br Med J 3, 759-763.<br />

McCord, J. M., and Day, E. D., Jr. (1978). Superoxide-dependent production <strong>of</strong><br />

hydroxyl radical catalyzed by iron-EDTA complex. FEBS Lett 86, 139-142.<br />

McCune, W. J., Golbus, J., Zeldes, W., Bohlke, P., Dunne, R., and Fox, D. (1988).<br />

Clinical and immunologic effects <strong>of</strong> monthly administration <strong>of</strong> intravenous<br />

cyclophosphamide in severe systemic lupus ery<strong>the</strong>matous. N Engl J Med 318,<br />

1423-1431.<br />

McDonagh, B. J., and Martin, J. (1970). Paraquat poisoning in children. Arch Dis Child<br />

45, 425-427.<br />

McKeag, D., Maini, R., and Taylor, H. R. (2002). The ocular surface toxicity <strong>of</strong><br />

<strong>paraquat</strong>. Br J Ophthalmol 86, 350-351.<br />

McNemar, M. D., Gorman, J. A., and Buckley, H. R. (2001). Isolation <strong>of</strong> a gene<br />

encoding a putative polyamine transporter from Candida albicans, GPT1. Yeast<br />

18, 555-561.<br />

Mees, G. C. (1960). Experiments on <strong>the</strong> herbicidal action <strong>of</strong> 1,1'-ethylene-<br />

2,2'dipyridylium dibromide. Ann Appl Biol 48, 601-612.<br />

Mehani, S. (1972). The toxic effect <strong>of</strong> <strong>paraquat</strong> in rabbits and rats. Ain Shams Med J,<br />

599-601.<br />

Meredith, T. J., and Vale, J. A. (1987). Treatment <strong>of</strong> <strong>paraquat</strong> poisoning in man:<br />

methods to prevent absorption. Hum Toxicol 6, 49-55.<br />

Michaelis, L., and Hill, E. S. (1933). The viologen indicators. J Gen Physiol 16, 859.<br />

Miles, A. T., Hawksworth, G. M., Beattie, J. H., and Rodilla, V. (2000). Induction,<br />

regulation, degradation, and biological significance <strong>of</strong> mammalian<br />

metallothioneins. Crit Rev Biochem Mol Biol 35, 35-70.<br />

Modee, J., Ivemark, B. I., and Robertson, B. (1972). Ultrastructure <strong>of</strong> <strong>the</strong> alveolar wall<br />

in experimental <strong>paraquat</strong> poisoning. Acta Pathol Microbiol Scand 80, 54-60.<br />

Mohammadi-Karakani, A., Ghazi-Khansari, M., and Sotoudeh, M. (2006). Lisinopril<br />

ameliorates <strong>paraquat</strong>-<strong>induced</strong> lung fibrosis. Clin Chim Acta 367, 170-174.<br />

Murphy, P. G., Myers, D. S., Davies, M. J., Webster, N. R., and Jones, J. G. (1992). The<br />

antioxidant potential <strong>of</strong> prop<strong>of</strong>ol (2,6-diisopropylphenol). Br J Anaesth 68, 613-<br />

618.<br />

Murray, R. E., and Gibson, J. E. (1972). A comparative study <strong>of</strong> <strong>paraquat</strong> <strong>into</strong>xication<br />

in rats, guinea pigs and monkeys. Exp Mol Pathol 17, 317-325.<br />

Nagao, M., Saitoh, H., Zhang, W. D., Iseki, K., Yamada, Y., Takatori, T., and<br />

Miyazaki, K. (1993a). Transport characteristics <strong>of</strong> <strong>paraquat</strong> across rat intestinal<br />

brush-border membrane. Arch Toxicol 67, 262-267.<br />

240


________________________________________________________Part IV – References<br />

Nagao, M., Takatori, T., Wu, B., Terazawa, K., Gotouda, H., and Akabane, H. (1989).<br />

Immuno<strong>the</strong>rapy for <strong>the</strong> treatment <strong>of</strong> acute <strong>paraquat</strong> poisoning. Hum Toxicol 8,<br />

121-123.<br />

Nagao, M., Zhang, W. D., Itakura, Y., Kobayashi, M., Yamada, Y., Yagi, K., Oono, T.,<br />

and Takatori, T. (1993b). Immunohistochemical localization and dynamics <strong>of</strong><br />

<strong>paraquat</strong> in <strong>the</strong> stomach and esophagus <strong>of</strong> rats. Int J Legal Med 106, 142-144.<br />

Nagao, M., Zhang, W. D., Itakura, Y., Kobayashi, M., Yamada, Y., Yagi, K., Oono, T.,<br />

and Takatori, T. (1993c). Immunohistochemical localization <strong>of</strong> <strong>paraquat</strong> in skin<br />

and eye <strong>of</strong> rat. Nippon Hoigaku Zasshi 47, 202-206.<br />

Nagao, M., Zhang, W. D., Takatori, T., Itakura, Y., Yamada, Y., Iwase, H., Oono, T.,<br />

and Iwadate, K. (1994). Identification and dynamics <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> bone<br />

marrow, thymus and spleen in rats using immunohistochemical techniques.<br />

Nippon Hoigaku Zasshi 48, 166-168.<br />

Nagata, M. (2005). Inflammatory cells and oxygen radicals. Curr Drug Targets Inflamm<br />

Allergy 4, 503-504.<br />

Nagata, T., Kono, I., Masoaka, T., and Akahori, F. (1992). Acute toxicological studies<br />

on <strong>paraquat</strong>: pathological findings in beagle dogs following single subcutaneous<br />

injections. Vet Hum Toxicol 34, 105-111.<br />

Nagi, A. H. (1970). Paraquat and adrenal cortical necrosis. Br Med J 2, 669.<br />

Nakagawa, I., Suzuki, M., Imura, N., and Naganuma, A. (1995). Enhancement <strong>of</strong><br />

<strong>paraquat</strong> toxicity by glutathione depletion in mice in vivo and in vitro. J Toxicol<br />

Sci 20, 557-564.<br />

Nakagawa, I., Suzuki, M., Imura, N., and Naganuma, A. (1998). Involvement <strong>of</strong><br />

oxidative stress in <strong>paraquat</strong>-<strong>induced</strong> metallothionein syn<strong>the</strong>sis under glutathione<br />

depletion. Free Radic Biol Med 24, 1390-1395.<br />

Nakamura, T., Ushiyama, C., Shimada, N., Hayashi, K., Ebihara, I., Suzaki, M., and<br />

Koide, H. (2001). Changes in concentrations <strong>of</strong> type IV collagen and tissue<br />

inhibitor <strong>of</strong> metalloproteinase-1 in patients with <strong>paraquat</strong> poisoning. J Appl<br />

Toxicol 21, 445-447.<br />

Nemery, B., Smith, L. L., and Aldridge, W. N. (1987). Putrescine and 5hydroxytryptamine<br />

accumulation in rat lung slices: cellular localization and<br />

responses to cell-specific lung injury. Toxicol Appl Pharmacol 91, 107-120.<br />

Nemery, B., and van Klaveren, R. J. (1995). NO wonder <strong>paraquat</strong> is toxic. Hum Exp<br />

Toxicol 14, 308-309.<br />

Nemery, B., Vanlommel, S., Verbeken, E. K., Lauweryns, J. M., and Demedts, M.<br />

(1992). Lung injury <strong>induced</strong> by <strong>paraquat</strong>, hyperoxia and cobalt chloride: effects<br />

<strong>of</strong> ambroxol. Pulm Pharmacol Ther 5, 53-60.<br />

Newhouse, M., McEvoy, D., and Rosenthal, D. (1978). Percutaneous <strong>paraquat</strong><br />

absorption. An association with cutaneous lesions and respiratory failure. Arch<br />

Dermatol 114, 1516-1519.<br />

Nirei, M., Hayasaka, S., Nagata, M., Tamai, A., and Tawara, T. (1993). Ocular injury<br />

caused by Preeglox-L, a herbicide containing <strong>paraquat</strong>, diquat and surfactants.<br />

Jpn J Ophthalmol 37, 43-46.<br />

Noguchi, N., Misawa, S., Tsuchiya, S., Yamamoto, H., and Naito, H. (1985). Cardiorespiratory<br />

effects <strong>of</strong> <strong>paraquat</strong> with and without emetics on Wistar rats. Vet Hum<br />

Toxicol 27, 508-510.<br />

O'Sullivan, M. C., Golding, B. T., Smith, L. L., and Wyatt, I. (1991). Molecular features<br />

necessary for <strong>the</strong> uptake <strong>of</strong> diamines and related compounds by <strong>the</strong> polyamine<br />

receptor <strong>of</strong> rat lung slices. Biochem. Pharmacol 41, 1839-1848.<br />

241


Part IV – References____________________________________________________________________<br />

Okonek, S., Baldamus, C. A., H<strong>of</strong>mann, A., Schuster, C. J., Bechstein, P. B., and Zoller,<br />

B. (1979). Two survivors <strong>of</strong> severe <strong>paraquat</strong> <strong>into</strong>xication by "continuous<br />

hemoperfusion". Klin Wochenschr 57, 957-959.<br />

Okonek, S., H<strong>of</strong>mann, A., and Henningsen, B. (1976). Efficacy <strong>of</strong> gut lavage,<br />

hemodialysis, and hemoperfusion in <strong>the</strong> <strong>the</strong>rapy <strong>of</strong> <strong>paraquat</strong> or diquat<br />

<strong>into</strong>xication. Arch Toxicol 36, 43-51.<br />

Okonek, S., Setyadharma, H., Borchert, A., and Krienke, E. G. (1982a). Activated<br />

charcoal is as effective as fuller's earth or bentonite in <strong>paraquat</strong> poisoning. Klin<br />

Wochenschr 60, 207-210.<br />

Okonek, S., Weilemann, L. S., Majdandzic, J., Setyadharma, H., Reinecke, H. J.,<br />

Baldamus, C. A., Lohmann, J., Bonzel, K. E., and Thon, T. (1982b). Successful<br />

treatment <strong>of</strong> <strong>paraquat</strong> poisoning: activated charcoal per os and "continuous<br />

hemoperfusion". J Toxicol Clin Toxicol 19, 807-819.<br />

Okonek, S., Wronski, R., Niedermayer, W., Okonek, M., and Lamer, A. (1983). Near<br />

fatal percutaneous <strong>paraquat</strong> poisoning. Klin Wochenschr 61, 655-659.<br />

Oliveira, M. V., Albuquerque, J. A., Paixao, A. D., Guedes, L. S., and Cabral, A. M.<br />

(2005). High blood pressure is one <strong>of</strong> <strong>the</strong> symptoms <strong>of</strong> <strong>paraquat</strong>-<strong>induced</strong> toxicity<br />

in rats. Arch Toxicol 79, 515-518.<br />

Omaye, S. T., Reddy, K. A., and Cross, C. E. (1978). Enhanced lung toxicity <strong>of</strong><br />

<strong>paraquat</strong> in selenium-deficient rats. Toxicol Appl Pharmacol 43, 237-247<br />

Ong, M. L., and Glew, S. (1989). Paraquat poisoning: Per vagina. Postgrad Med J 65,<br />

835-836.<br />

Onyon, L. J., and Volans, G. N. (1987). The epidemiology and prevention <strong>of</strong> <strong>paraquat</strong><br />

poisoning. Hum Toxicol 6, 19-29.<br />

Orme, S., and Kegley, S. PAN Pesticide Database and Pesticide Action Network North<br />

America (San Francisco, CA 2006),<br />

http://www.pesticideinfo.org/List_NTPStudies.jsp?Rec_Id=PC33358.<br />

Pan, T., Nielsen, L. D., Allen, M. J., Shannon, K. M., Shannon, J. M., Selman, M., and<br />

Mason, R. J. (2002). Serum SP-D is a marker <strong>of</strong> lung injury in rats. Am J Physiol<br />

Lung Cell Mol Physiol 282, L824-L832.<br />

Papiris, S. A., Maniati, M. A., Kyriakidis, V., and Constantopoulos, S. H. (1995).<br />

Pulmonary damage due to <strong>paraquat</strong> poisoning through skin absorption.<br />

Respiration 62, 101-103.<br />

Parkins, C. S., and Fowler, J. F. (1985). A cautionary note on <strong>the</strong> resolution <strong>of</strong> <strong>paraquat</strong><br />

lung damage after radio<strong>the</strong>rapy. Br J Radiol 58, 1137-1140.<br />

Patterson, C. E., and Rhoades, R. A. (1988). Protective role <strong>of</strong> sulfhydryl reagents in<br />

oxidant lung injury. Exp Lung Res 14 Suppl, 1005-1019.<br />

Perriens, J. H., Benimadho, S., Kiauw, I. L., Wisse, J., and Chee, H. (1992). High-dose<br />

cyclophosphamide and dexamethasone in <strong>paraquat</strong> poisoning: a prospective<br />

study. Hum Exp Toxicol 11, 129-134.<br />

Peters, J. M. (1967). Effects <strong>of</strong> absolute starvation and refeeding on <strong>organ</strong> weights and<br />

water contents <strong>of</strong> albino rats. Growth 31, 191-203.<br />

Petty, M., Grisar, J. M., Dow, J., and De Jong, W. (1990). Effects <strong>of</strong> an alphatocopherol<br />

analogue on myocardial ischaemia and reperfusion injury in rats. Eur<br />

J Pharmacol 179, 241-242.<br />

Phillips, M. (1992). Breath tests in medicine. Sci Am 267, 74-79.<br />

Pietra, G. G. (1984). New insights <strong>into</strong> <strong>mechanisms</strong> <strong>of</strong> pulmonary edema. Lab Invest 51,<br />

489-494.<br />

Pond, S. (1990). Manifestations and management <strong>of</strong> <strong>paraquat</strong> poisoning. Med J Aust<br />

152, 256-259.<br />

242


________________________________________________________Part IV – References<br />

Pozzi, E., Luisetti, M., Spialtini, L., Coccia, P., Rossi, A., Donnini, M., Cetta, G., and<br />

Salmona, M. (1989). Relationship between changes in alveolar surfactant levels<br />

and lung defence <strong>mechanisms</strong>. Respiration 55, 53-59.<br />

Proudfoot, A. T. (1983). Toxicity <strong>of</strong> salicylates. Am J Med 75, 99-103.<br />

Proudfoot, A. T. (1995). Predictive value <strong>of</strong> early plasma <strong>paraquat</strong> concentrations. In<br />

Paraquat Poisoning: Mechanisms, Prevention, Treatment (C. Bismuth, and A.<br />

H. Hall, Eds.), pp. 275-283. Marcel Dekker, New York.<br />

Proudfoot, A. T. (1999). Paraquat: <strong>mechanisms</strong> <strong>of</strong> toxicity, clinical features and<br />

management. Clin. Toxicol 37, 360-361.<br />

Proudfoot, A. T., Prescott, L. F., and Jarvie, D. R. (1987). Haemodialysis for <strong>paraquat</strong><br />

poisoning. Hum Toxicol 6, 69-74.<br />

Proudfoot, A. T., Stewart, M. S., Levitt, T., and Widdop, B. (1979). Paraquat poisoning:<br />

significance <strong>of</strong> plasma-<strong>paraquat</strong> concentrations. Lancet 18, 330-332.<br />

Purser, D. A., and Rose, M. S. (1979). The toxicity and renal handling <strong>of</strong> <strong>paraquat</strong> in<br />

cynomolgus monkeys. Toxicology 15, 31-41.<br />

Ranjbar, A., Pasalar, P., Sedighi, A., and Abdollahi, M. (2002). Induction <strong>of</strong> oxidative<br />

stress in <strong>paraquat</strong> formulating workers. Toxicol Lett 131, 191-194.<br />

Rannels, D. E., Kameji, R., Pegg, A. E., and Rannels, S. R. (1989). Spermidine uptake<br />

by type II pneumocytes: interactions <strong>of</strong> amine uptake pathways. Am J Physiol<br />

Lung Cell Mol Physiol 257, L346-L353.<br />

Rannels, D. E., Pegg, A. E., Clark, R. S., and Addison, J. L. (1985). Interaction <strong>of</strong><br />

<strong>paraquat</strong> and amine uptake by rat lungs perfused in situ. Am J Physiol<br />

Endocrinol Metab 249, E506-E513.<br />

Redetzki, H. M., Wood, C. D., and Grafton, W. D. (1980). Vitamin E and <strong>paraquat</strong><br />

poisoning. Vet Hum Toxicol 22, 395-397.<br />

Reif, R. M., and Lewinsohn, G. (1983). Paraquat myocarditis and adrenal cortical<br />

necrosis. J Forensic Sci Soc 28, 505-509.<br />

Rhodes, M. L., Zavala, D. C., and Brown, D. (1976). Hypoxic protection in <strong>paraquat</strong><br />

poisoning. Lab Invest 35, 496-500.<br />

Ricciardolo, F. L., Di Stefano, A., Sabatini, F., and Folkerts, G. (2006). Reactive<br />

nitrogen species in <strong>the</strong> respiratory tract. Eur J Pharmacol 533, 240-252.<br />

Richmond, R., and Halliwell, B. (1982). Formation <strong>of</strong> hydroxyl radicals from <strong>the</strong><br />

<strong>paraquat</strong> radical cation, demonstrated by a highly specific gas chromatographic<br />

technique. <strong>the</strong> role <strong>of</strong> superoxide radical anion, hydrogen peroxide, and<br />

glutathione reductase. J Inorg Biochem 17, 95-107.<br />

Roberts, T. R., Dyson, J. S., and Lane, M. C. (2002). Deactivation <strong>of</strong> <strong>the</strong> biological<br />

activity <strong>of</strong> <strong>paraquat</strong> in <strong>the</strong> soil environment: a review <strong>of</strong> long-term<br />

environmental fate. J Agric Food Chem 50, 3623-3631.<br />

Robertson, B., Enhorning, G., Ivemark, B., Malmqvist, E., and Modee, J. (1970).<br />

Paraquat-<strong>induced</strong> derangement <strong>of</strong> pulmonary surfactant in <strong>the</strong> rat. Acta Paediatr<br />

Scand Suppl 206, 37-39.<br />

Rose, M. S., Lock, E. A., Smith, L. L., and Wyatt, I. (1976a). Paraquat accumulation:<br />

tissue and species specificity. Biochem Pharmacol 25, 419-423.<br />

Rose, M. S., and Smith, L. L. (1977). Tissue uptake <strong>of</strong> <strong>paraquat</strong> and diquat. Gen<br />

Pharmacol 8, 173-176.<br />

Rose, M. S., Smith, L. L., and Wyatt, I. (1974). Evidence for energy-dependent<br />

accumulation <strong>of</strong> <strong>paraquat</strong> <strong>into</strong> rat lung. Nature 252, 314-315.<br />

Rose, M. S., Smith, L. L., and Wyatt, I. (1976b). The relevance <strong>of</strong> pentose phosphate<br />

pathway stimulation in rat lung to <strong>the</strong> mechanism <strong>of</strong> <strong>paraquat</strong> toxicity. Biochem<br />

Pharmacol 25, 1763-1767.<br />

243


Part IV – References____________________________________________________________________<br />

Ross, J. H., and Krieger, R. I. (1981). Structure-activity correlations <strong>of</strong> amines<br />

inhibiting active uptake <strong>of</strong> <strong>paraquat</strong> (methyl viologen) <strong>into</strong> rat lung slices.<br />

Toxicol Appl Pharmacol 59, 238-249.<br />

Russell, J. H., and Wallwork, S. C. (1972). The crystal structures <strong>of</strong> <strong>the</strong> dichloride and<br />

isomorphous dibromide and diiodide <strong>of</strong> <strong>the</strong> N,N'-dimethyl-4,4'-bipyridylium ion.<br />

Acta Cryst B 28, 1527-1533.<br />

Russell, L. A., Stone, B. E., and Rooney, P. A. (1981). Paraquat poisoning: toxicologic<br />

and pathologic findings in three fatal cases. Clin Toxicol 18, 915-928.<br />

Sagar, G. R. (1987). Uses and usefulness <strong>of</strong> <strong>paraquat</strong>. Hum Toxicol 6, 7-11.<br />

Sakai, M., Yamagami, K., Kawamoto, K., and Tanaka, T. (1993). Tungsten modulates<br />

<strong>the</strong> toxicity <strong>of</strong> <strong>paraquat</strong> for epi<strong>the</strong>lial cells. Hum Cell 6, 287-293.<br />

Sakai, M., Yamagami, K., Kitazawa, Y., Takeyama, N., and Tanaka, T. (1995).<br />

Xanthine oxidase mediates <strong>paraquat</strong>-<strong>induced</strong> toxicity on cultured endo<strong>the</strong>lial<br />

cell. Pharmaco Toxicol 77, 36-40.<br />

Salmona, M., Donnini, M., Perin, L., Diomede, L., Romano, M., Marini, M. G.,<br />

Tacconi, M. T., and Luisetti, M. (1992). A novel pharmacological approach for<br />

<strong>paraquat</strong> poisoning in rat and A549 cell line using ambroxol, a lung surfactant<br />

syn<strong>the</strong>sis inducer. Food Chem Toxicol 30, 789-794.<br />

Salovsky, P., and Shopova, V. (1993). Synergic lung changes in rats receiving<br />

combined exposure to <strong>paraquat</strong> and ionizing radiation. Environ Res 60, 44-54.<br />

Samman, P. D., and Johnston, E. N. (1969). Nail damage associated with handling <strong>of</strong><br />

<strong>paraquat</strong> and diquat. Br Med J 1, 818-819.<br />

Sato, M., Apostolova, M. D., Hamaya, M., Yamaki, J., Choo, K. H. A., Michalska, A.<br />

E., Kodama, N., and Tohyama, C. (1996). Susceptibility <strong>of</strong> metallothionein-null<br />

mice to <strong>paraquat</strong>. Environ Toxicol. Pharmacol. 1, 221-225.<br />

Saunders, N. A., Ilett, K. F., and Minchin, R. F. (1989). Pulmonary alveolar<br />

macrophages express a polyamine transport system. J Cell Physiol 139, 624-<br />

631.<br />

Saunders, N. R., Alpert, H. M., and Cooper, J. D. (1985). Sequential bilateral lung<br />

transplantation for <strong>paraquat</strong> poisoning. A case report. J Thorac Cardiovasc Surg<br />

89, 734-742.<br />

Sawada, Y., Yamamoto, I., Hirokane, T., Nagai, Y., Satoh, Y., and Ueyama, M. (1988).<br />

Severity index <strong>of</strong> <strong>paraquat</strong> poisoning. Lancet 1, 1333.<br />

Scherrmann, J. M. (1995). Analytical procedures and predictive value <strong>of</strong> late plasma<br />

and urine <strong>paraquat</strong> concentrations. In Paraquat poisoning: <strong>mechanisms</strong>,<br />

prevention, treatment (C. Bismuth, and A. H. Hall, Eds.), pp. 285-296. Marcel<br />

Dekker, New York.<br />

Scherrmann, J. M., Houze, P., Bismuth, C., and Bourdon, R. (1987). Prognostic value <strong>of</strong><br />

plasma and urine <strong>paraquat</strong> concentration. Hum Toxicol 6, 91-93.<br />

Schoenberger, C. I., Rennard, S. I., Bitterman, P. B., Fukuda, Y., Ferrans, V. J., and<br />

Crystal, R. G. (1984). Paraquat-<strong>induced</strong> pulmonary fibrosis. Role <strong>of</strong> <strong>the</strong><br />

alveolitis in modulating <strong>the</strong> development <strong>of</strong> fibrosis. Am Rev Respir Dis 129,<br />

168-173.<br />

Schvartsman, S., Zyngier, S., and Schvartsman, C. (1984). Ascorbic acid and rib<strong>of</strong>lavin<br />

in <strong>the</strong> treatment <strong>of</strong> acute <strong>into</strong>xication by <strong>paraquat</strong>. Vet Hum Toxicol 26, 473-475.<br />

Schweich, M. D., Lison, D., and Lauwerys, R. (1994). Assessment <strong>of</strong> lipid peroxidation<br />

associated with lung damage <strong>induced</strong> by oxidative stress. In vivo and in vitro<br />

studies. Biochem Pharmacol 47, 1395-1400.<br />

Seidenfeld, J. J., Wyc<strong>of</strong>f, D., Zavala, D. C., and Richerson, H. B. (1978). Paraquat lung<br />

injury in rabbits. Br J Ind Med 35, 245-257.<br />

244


________________________________________________________Part IV – References<br />

Senanayake, N., Gurunathan, G., Hart, T. B., Amerasinghe, P., Babapulle, M., Ellapola,<br />

S. B., Udupihille, M., and Basanayake, V. (1993). An epidemiological study <strong>of</strong><br />

<strong>the</strong> health <strong>of</strong> Sri Lankan tea plantation workers associated with long term<br />

exposure to <strong>paraquat</strong>. Br J Ind Med 50, 257-263.<br />

Services, U. D. o. H. a. H. (1988). NIOSH recommendations for occupational safety and<br />

health standards, 1988. MMWR Morb Mortal Wkly Rep 37(suppl 7), 1-29.<br />

Shahar, E., Barzilay, Z., and Aladjem, M. (1980). Paraquat poisoning in a child: vitamin<br />

E in amelioration <strong>of</strong> lung injury. Arch Dis Child 55, 830-831.<br />

Sharp, C. W., Ottolenghi, A., and Posner, H. S. (1972). Correlation <strong>of</strong> <strong>paraquat</strong> toxicity<br />

with tissue concentrations and weight loss <strong>of</strong> <strong>the</strong> rat. Toxicol Appl Pharmacol<br />

22, 241-251.<br />

Shirahama, M., Sakemi, T., Osato, S., Sanai, T., Rikitake, O., and Wada, S. (1987).<br />

Recovery after radio<strong>the</strong>rapy from severe interstitial pneumonia due to <strong>paraquat</strong><br />

poisoning. Jpn J Med 26, 385-387.<br />

Shu, H., Talcott, R. E., Rice, S. A., and Wei, E. T. (1979). Lipid peroxidation and<br />

<strong>paraquat</strong> toxicity. Biochem Pharmacol 28, 327-331.<br />

Siddik, Z. H., Drew, R., and Gram, T. E. (1979). The effect <strong>of</strong> chlorpromazine on <strong>the</strong><br />

uptake and efflux <strong>of</strong> <strong>paraquat</strong> in rat lung slices. Toxicol Appl Pharmacol 50,<br />

443-450.<br />

Silva, M. F., and Saldiva, P. H. (1998). Paraquat poisoning: an experimental model <strong>of</strong><br />

dose-dependent acute lung injury due to surfactant dysfunction. Braz J Med Biol<br />

Res 31, 445-450.<br />

Sinow, J., and Wei, E. (1973). Ocular toxicity <strong>of</strong> <strong>paraquat</strong>. Bull Environ Contam<br />

Toxicol 9, 163-168.<br />

Situnayake, R. D., Crump, B. J., Thurnham, D. I., Davies, J. A., and Davis, M. (1987).<br />

Evidence for lipid peroxidation in man following <strong>paraquat</strong> ingestion. Hum<br />

Toxicol 6, 94-98.<br />

Slade, P. (1965). Photochemical degradation <strong>of</strong> <strong>paraquat</strong>. Nature (Lond.) 207, 515.<br />

Slade, P. (1966). The fate <strong>of</strong> <strong>paraquat</strong> applied to plants. Weed Res 6, 158-167.<br />

Smith, E. A., and Mayfield, C. I. (1978). Paraquat: Determination, degradation, and<br />

mobility in soil. Water, Air, and Soil Pollution 9, 439-452.<br />

Smith, E. A., and Oehme, F. W. (1991). A review <strong>of</strong> selected herbicides and <strong>the</strong>ir<br />

toxicities. Vet Hum Toxicol 33, 596-608.<br />

Smith, I., White, P. F., Nathanson, M., and Gouldson, R. (1994). Prop<strong>of</strong>ol. An update<br />

on its clinical use. Anes<strong>the</strong>siology 81, 1005-1043.<br />

Smith, I. C., and Schneider, W. G. (1961). The proton magnetic resonance spectrum and<br />

<strong>the</strong> charge distribution <strong>of</strong> <strong>the</strong> pyridinium ion. Can J Chem 39, 1158-1161.<br />

Smith, J. G. (1988a). Paraquat poisoning by skin absorption: a review. Hum Toxicol 7,<br />

15-19.<br />

Smith, L. J., Anderson, J., and Shamsuddin, M. (1992). Glutathione localization and<br />

distribution after intratracheal instillation. Am Rev Respir Dis 145, 153-159.<br />

Smith, L. L. (1982). The identification <strong>of</strong> an accumulation system for diamines and<br />

polyamines <strong>into</strong> <strong>the</strong> lung and its relevance to <strong>paraquat</strong> toxicity. Arch Toxicol<br />

Suppl 5, 1-14.<br />

Smith, L. L. (1987). Mechanism <strong>of</strong> <strong>paraquat</strong> toxicity in lung and its relevance to<br />

treatment. Hum Toxicol 6, 31-36.<br />

Smith, L. L. (1988b). The toxicity <strong>of</strong> <strong>paraquat</strong>. Adverse Drug React Acute Poisoning<br />

Rev 7, 1-17.<br />

Smith, L. L., and Nemery, B. (1986). The lung as a target <strong>organ</strong> for toxicity. In Target<br />

Organ Toxicity (G. M. Cohen, Ed.), pp. 45-80. CRC Press, Boca Raton, FL.<br />

245


Part IV – References____________________________________________________________________<br />

Smith, L. L., and Nemery, B. (1992). Cellular specific toxicity in <strong>the</strong> lung. In Selectivity<br />

and Molecular Mechanisms <strong>of</strong> Toxicity (F. D. Matteis, and E. A. Lock, Eds.),<br />

pp. 3-26. Macmillan, London.<br />

Smith, L. L., Rose, M. S., and Wyatt, I. (1978). The pathology and biochemistry <strong>of</strong><br />

<strong>paraquat</strong>. Ciba Found Symp 65, 321-341.<br />

Smith, L. L., Wright, A., Wyatt, I., and Rose, M. S. (1974a). Effective treatment for<br />

<strong>paraquat</strong> poisoning in rats and its relevance to treatment <strong>of</strong> <strong>paraquat</strong> poisoning in<br />

man. Br Med J 4, 569-571.<br />

Smith, L. L., and Wyatt, I. (1981). The accumulation <strong>of</strong> putrescine <strong>into</strong> slices <strong>of</strong> rat lung<br />

and brain and its relationship to <strong>the</strong> accumulation <strong>of</strong> <strong>paraquat</strong>. Biochem<br />

Pharmacol 30, 1053-1058.<br />

Smith, P., and Heath, D. (1974). The ultrastructure and time sequence <strong>of</strong> <strong>the</strong> early<br />

stages <strong>of</strong> <strong>paraquat</strong> lung in rats. J Pathol 114, 177-184.<br />

Smith, P., and Heath, D. (1976). Paraquat. CRC Crit Rev Toxicol 4, 411-445.<br />

Smith, P., Heath, D., and Kay, J. M. (1974b). The pathogenesis and structure <strong>of</strong><br />

<strong>paraquat</strong>-<strong>induced</strong> pulmonary fibrosis in rats. J Pathol 114, 57-67.<br />

So, K. L., de Buijzer, E., Gommers, D., Kaisers, U., van Genderen, P. J., and<br />

Lachmann, B. (1998). Surfactant <strong>the</strong>rapy restores gas exchange in lung injury<br />

due to <strong>paraquat</strong> <strong>into</strong>xication in rats. Eur Respir J 12, 284-287.<br />

S<strong>of</strong>uni, T., Hatanaka, M., and Ishidate, M. (1988). Chromosomal aberrations and<br />

superoxide generating systems II. Effects <strong>of</strong> <strong>paraquat</strong> on Chinese hamster cells<br />

in vitro. Mutat Res 147, 273-274.<br />

Sorokin, S. P. (1970). The cells <strong>of</strong> <strong>the</strong> lungs. In Morphology <strong>of</strong> experimental<br />

Respiratory Carcinogenesis (P. Nettesheim, M G Hanna, and J. W. Dea<strong>the</strong>ridge,<br />

Eds.). US Atomic Energy Commission, D C Washington<br />

Stephens, D. S., Walker, D. H., Schaffner, W., Kaplowitz, L. G., Brashear, H. R.,<br />

Roberts, R., and Spickard , W. A., Jr. (1981). Pseudodiph<strong>the</strong>ria: prominent<br />

pharyngeal membrane associated with fatal <strong>paraquat</strong> ingestion. Ann Intern Med<br />

94, 202-204.<br />

Stratta, P., Mazzucco, G., Tetta, S. G. C., and Monga, G. (1988). Immune-mediated<br />

glomerulonephritis after exposure to <strong>paraquat</strong>. Nephron 48, 138-141.<br />

Sugihara, N., Suetsugu, T., and Furuno, K. (1995). High susceptibility to <strong>paraquat</strong>driven<br />

lipid peroxidation <strong>of</strong> cultured hepatocytes loaded with linolenic acid. J<br />

Pharmacol Exper Therapeut 274, 187-193.<br />

Sullivan, T. M., and Montgomery, M. R. (1983). The relationship between <strong>paraquat</strong><br />

accumulation and covalent binding in rat lung slices. Drug Metab Dispos 11,<br />

526-530.<br />

Sullivan, T. M., and Montgomery, M. R. (1986). Glucose ameliorates <strong>the</strong> depletion <strong>of</strong><br />

NADPH by <strong>paraquat</strong> in rat lung slices. Toxicology 41, 145-152.<br />

Summers, L. A. (1980). The bipyridilium herbicides. Academic Press, London, New<br />

York, Toronto, San Francisco.<br />

Suntres, Z. E., and Shek, P. N. (1995). Intratracheally administered liposomal alphatocopherol<br />

protects <strong>the</strong> lung against long-term toxic effects <strong>of</strong> <strong>paraquat</strong>. Biomed<br />

Environ Sci 8, 289-300.<br />

Suzuki, K., Takasu, N., Arita, S., Maenosono, A., Ishimatsu, S., Nishina, M., Tanaka,<br />

S., and Kohama, A. (1989). A new method for predicting <strong>the</strong> outcome and<br />

survival period in <strong>paraquat</strong> poisoning. Hum Exp Toxicol 8, 33-38.<br />

Suzuki, K., Takasu, N., Arita, S., Ueda, A., Okabe, T., Ishimatsu, S., Tanaka, S., and<br />

Kohama, A. (1991). Evaluation <strong>of</strong> severity indexes <strong>of</strong> patients with <strong>paraquat</strong><br />

poisoning. Hum Exp Toxicol 10, 21-23.<br />

246


________________________________________________________Part IV – References<br />

Sykes, B. I., Purchase, I. F., and Smith, L. L. (1977). Pulmonary ultrastructure after oral<br />

and intravenous dosage <strong>of</strong> <strong>paraquat</strong> to rats. J Pathol 121, 233-241.<br />

Tabata, N., Morita, M., Mimasaka, S., Funayama, M., Hagiwara, T., and Abe, M.<br />

(1999). Paraquat myopathy: report on two suicide cases. Forensic Sci Int 100,<br />

117-126.<br />

Tabei, K., Asano, Y., and Hosoda, S. (1982). Efficacy <strong>of</strong> charcoal hemoperfusion in<br />

<strong>paraquat</strong> poisoning. Artif Organs 6, 37-42.<br />

Takeyama, N., Tanaka, T., Yabuki, T., and Nakatani, T. (2004). The involvement <strong>of</strong><br />

p53 in <strong>paraquat</strong>-<strong>induced</strong> apoptosis in human lung epi<strong>the</strong>lial-like cells. Int J<br />

Toxicol 23, 33-40.<br />

Talbot, A. R., and Barnes, M. R. (1988). Radio<strong>the</strong>rapy for <strong>the</strong> treatment <strong>of</strong> pulmonary<br />

complications <strong>of</strong> <strong>paraquat</strong> poisoning. Hum. Toxicol 7, 325-332.<br />

Talbot, A. R., Fu, C. C., and Hsieh, M. F. (1988). Paraquat <strong>into</strong>xication during<br />

pregnancy: a report <strong>of</strong> 9 cases. Vet Hum Toxicol 30, 12-17.<br />

Teixeira, H., Proenca, P., Alvarenga, M., Oliveira, M., Marques, E. P., and Vieira, D. N.<br />

(2004). Pesticide <strong>into</strong>xications in <strong>the</strong> Centre <strong>of</strong> Portugal: three years analysis.<br />

Forensic Sci Int 143, 199-204.<br />

Tokunaga, I., Kubo, S., Mikasa, H., Suzuki, Y., and Morita, K. (1997). Determination<br />

<strong>of</strong> 8-hydroxy-deoxyguanosine formation in rat <strong>organ</strong>s: assessment <strong>of</strong> <strong>paraquat</strong>evoked<br />

oxidative DNA damage. Biochem Mol Biol Int 43, 73-77.<br />

Tompsett, S. L. (1970). Paraquat poisoning. Acta Pharmacol Toxicol (Copenh) 28, 346-<br />

358<br />

Troncy, E., Francoeur, M., and Blaise, G. (1997). Inhaled nitric oxide: clinical<br />

applications, indications, and toxicology. Can J Anaesth 44, 973-988.<br />

Tsatsakis, A. M., Perakis, K., and Koumantakis, E. (1996). Experience with acute<br />

<strong>paraquat</strong> poisoning in Crete. Vet Hum Toxicol 38, 113-117.<br />

Vale, J. A., Meredith, T. J., and Buckley, B. M. (1987). Paraquat poisoning: clinical<br />

features and immediate general management. Hum Toxicol 6, 41–47.<br />

van Asbeck, B. S., Hillen, F. C., Boonen, H. C., de Jong, Y., Dormans, J. A., van der<br />

Wal, N. A., Marx, J. J., and Sangster, B. (1989). Continuous intravenous<br />

infusion <strong>of</strong> deferoxamine reduces mortality by <strong>paraquat</strong> in vitamin E-deficient<br />

rats. Am Rev Respir Dis 139, 769-773.<br />

Van de Vyver, F. L., Giuliano, R. A., Paulus, G. J., Verpooten, G. A., Franke, J. P., De<br />

Zeeuw, R. A., Van Gaal, L. F., and De Broe, M. E. (1985). Hemoperfusionhemodialysis<br />

ineffective for <strong>paraquat</strong> removal in life-threatening poisoning?<br />

Clin Toxicol 23, 117-131.<br />

Van der Wal, N. A., Smith, L. L., van Oirschot, J. F., and van Asbeck, B. S. (1992).<br />

Effect <strong>of</strong> iron chelators on <strong>paraquat</strong> toxicity in rats and alveolar type II cells. Am<br />

Rev Respir Dis 145, 180-186.<br />

Vandenbogaerde, J., Schelstraete, J., Colardyn, F., and Heyndrickx, A. (1984). Paraquat<br />

poisoning. Forensic Sci Int 26, 103-114.<br />

Vaziri, N. D., Ness, R. L., Fairshter, R. D., Smith, W. R., and Rosen, S. M. (1979).<br />

Nephrotoxicity <strong>of</strong> <strong>paraquat</strong> in man. Arch Intern Med 139, 172-174.<br />

Vijeyaratnam, G. S., and Corrin, B. (1971). Experimental <strong>paraquat</strong> poisoning: a<br />

histological and electron-optical study <strong>of</strong> <strong>the</strong> changes in <strong>the</strong> lung. J Pathol 103,<br />

123-129.<br />

Vile, G. F., and Winterbourn, C. C. (1988). Microsomal reduction <strong>of</strong> low-molecularweight<br />

Fe3+ chelates and ferritin: enhancement by adriamycin, <strong>paraquat</strong>,<br />

menadione, and anthraquinone 2-sulfonate and inhibition by oxygen. Arch<br />

Biochem Biophys 267, 606-613.<br />

247


Part IV – References____________________________________________________________________<br />

Waddell, W. J., and Marlowe, C. (1980). Tissue and cellular disposition <strong>of</strong> <strong>paraquat</strong> in<br />

mice. Toxicol Appl Pharmacol 56, 147-150.<br />

Waintrub, M. L., Terada, L. S., Beehler, C. J., Anderson, B. O., Leff, J. A., and Repine,<br />

J. E. (1990). Xanthine oxidase is increased and contributes to <strong>paraquat</strong>-<strong>induced</strong><br />

acute lung injury. J Appl Physiol 68, 1755-1757.<br />

Walker, M., Dugard, P. H., and Scott, R. C. (1983). Absorption through human and<br />

laboratory animal skins: In vitro comparisons. Acta Pharm Suec 20, 52-53.<br />

Wasserman, B., and Block, E. R. (1978). Prevention <strong>of</strong> acute <strong>paraquat</strong> toxicity in rats by<br />

superoxide dismutase. Aviat Space Environ Med 49, 805-809.<br />

Webb, D. B., and Leopold, J. D. (1983). Vasodilation and rehydration in <strong>paraquat</strong><br />

poisoning. Hum Toxicol 2, 531-534.<br />

Webb, D. B., Williams, M. V., Davies, B. H., and James, K. W. (1984). Resolution after<br />

radio<strong>the</strong>rapy <strong>of</strong> severe pulmonary damage due to <strong>paraquat</strong> poisoning. Br Med J<br />

(Clin Res Ed) 288, 1259-1260.<br />

Wegener, T., Sandhagen, B., Chan, K. W., and Saldeen, T. (1988). N-Acetylcysteine in<br />

<strong>paraquat</strong> toxicity: toxicological and histological evaluation in rats. Ups J Med<br />

Sci 93, 81-89.<br />

Weidel, M., and Rosso, M. (1882). Studien uber das Pyridin. Monatsschr Chem. 3, 850.<br />

Wesseling, C., Hogstedt, C., Picado, A., and Johansson, L. (1997). Unintentional fatal<br />

<strong>paraquat</strong> poisonings among agricultural workers in Costa Rica: report <strong>of</strong> 15<br />

cases. Am J Ind Med 32, 433-441.<br />

Wesseling, C., van Wendel de Joode, B., Ruepert, C., Leon, C., Monge, P., Hermosillo,<br />

H., and Partanen, T. J. (2001). Paraquat in developing countries. Int J Occup<br />

Environ Health 7, 275-286.<br />

Wester, R., Maibach, H. I., Bucks, D. A., and Aufrere, M. B. (1984). In vivo<br />

percutaneous absorption <strong>of</strong> <strong>paraquat</strong> from hand, leg, and forearm <strong>of</strong> humans. J<br />

Toxicol Environ Health A 14, 759-762.<br />

White, B. G. (1969). Bipyridylium Quaternary Salts and Related Compounds. III. Weak<br />

Intermolecular Charge-transfer Complexes <strong>of</strong> Biological Interest, occurring in<br />

Solution and involving <strong>paraquat</strong>. Trans Faraday Soc 65, 2000-2015.<br />

Widdop, B. (1976). Detection <strong>of</strong> <strong>paraquat</strong> in urine. Br Med J 2, 1135.<br />

Williams, P. S., Hendy, M. S., and Ackrill, P. (1984). Early management <strong>of</strong> <strong>paraquat</strong><br />

poisoning. Lancet 1, 627.<br />

Winchester, J. F. (2002). Dialysis and Hemoperfusion in Poisoning. Adv Ren Replace<br />

Ther 9, 26-30.<br />

Winterbourn, C. C. (1981). Production <strong>of</strong> hydroxyl radicals from <strong>paraquat</strong> radicals and<br />

H2O2. FEBS Lett 128, 339-342.<br />

Witschi, H., Haschek, W. M., Meyer, K. R., Ullrich, R. L., and Dalbey, W. E. (1980). A<br />

pathogenetic mechanism in lung fibrosis. Chest 78, 395-399.<br />

Witschi, H., Kacew, S., Hirai, K. I., and Cote, M. G. (1977). In vivo oxidation <strong>of</strong><br />

reduced nicotinamide-adenine dinucleotide phosphate by <strong>paraquat</strong> and diquat in<br />

rat lung. Chem Biol Interact 19, 143-160.<br />

Wohlfahrt, D. J. (1982). Fatal <strong>paraquat</strong> poisonings after skin absorption. Med J Aust 1,<br />

512-513.<br />

Wojeck, G. A., Price, J. F., Nigg, H. N., and Stamper, J. H. (1983). Worker exposure to<br />

<strong>paraquat</strong> and diquat. Arch Environ Contam Toxicol 12, 65-70.<br />

Wright, A. F., Green, T. P., Robson, R. T., Niewola, Z., Wyatt, I., and Smith, L. L.<br />

(1987). Specific polyclonal and monoclonal antibody prevents <strong>paraquat</strong><br />

accumulation <strong>into</strong> rat lung slices. Biochem Pharmacol 36, 1325-1331.<br />

248


________________________________________________________Part IV – References<br />

Wright, K. A., and Cain, R. B. (1970). Microbial degradation <strong>of</strong> 4-carboxy-1-methylpyridinium<br />

chloride, a photolytic product <strong>of</strong> <strong>paraquat</strong>. Biochem J 118, 52P-53P.<br />

Wyatt, I., Soames, A. R., Clay, M. F., and Smith, L. L. (1988). The accumulation and<br />

localisation <strong>of</strong> putrescine, spermidine, spermine and <strong>paraquat</strong> in <strong>the</strong> rat lung. In<br />

vitro and in vivo studies. Biochem Pharmacol 37, 1909-1918.<br />

Yamada, K., and Fukushima, T. (1993). Mechanism <strong>of</strong> cytotoxicity <strong>of</strong> <strong>paraquat</strong>. II.<br />

Organ specificity <strong>of</strong> <strong>paraquat</strong>-stimulated lipid peroxidation in <strong>the</strong> inner<br />

membrane <strong>of</strong> mitochondria. Exp Toxicol Pathol 45, 375-380.<br />

Yamamoto, H. (1993). Protection against <strong>paraquat</strong>-<strong>induced</strong> toxicity with sulfite or<br />

thiosulfate in mice. Toxicology 79, 37-43.<br />

Yamamoto, H. A., and Mohanan, P. V. (2001). Effects <strong>of</strong> melatonin on <strong>paraquat</strong> or<br />

ultraviolet light exposure-<strong>induced</strong> DNA damage. J Pineal Res 31, 308-313.<br />

Yamamoto, I., Saito, T., Harunari, N., Sato, Y., Kato, H., Nakagawa, Y., Inokuchi, S.,<br />

Sawada, Y., and Makuuchi, H. (2000). Correlating <strong>the</strong> severity <strong>of</strong> <strong>paraquat</strong><br />

poisoning with specific hemodynamic and oxygen metabolism variables. Crit<br />

Care Med 28, 1877-1883.<br />

Yamashita, M., Naito, H., and Takagi, S. (1987). The effectiveness <strong>of</strong> a cation resin<br />

(Kayexalate) as an adsorbent <strong>of</strong> <strong>paraquat</strong>: experimental and clinical studies.<br />

Hum Toxicol 6, 89-90.<br />

Yasaka, T., Ohya, I., Matsumoto, J., Shiramizu, T., and Sasaguri, Y. (1981).<br />

Acceleration <strong>of</strong> lipid peroxidation in human <strong>paraquat</strong> poisoning. Arch Intern<br />

Med 141, 1169-1171.<br />

Yasaka, T., Okudaira, K., Fujito, H., Matsumoto, J., Ohya, I., and Miyamoto, Y. (1986).<br />

Fur<strong>the</strong>r studies <strong>of</strong> lipid peroxidation in human <strong>paraquat</strong> poisoning. Arch Intern<br />

Med 146, 681-685.<br />

Yeh, S. T., Guo, H. R., Su, Y. S., Lin, H. J., Hou, C. C., Chen, H. M., Chang, M. C., and<br />

Wang, Y. J. (2006). Protective effects <strong>of</strong> N-acetylcysteine treatment post acute<br />

<strong>paraquat</strong> <strong>into</strong>xication in rats and in human lung epi<strong>the</strong>lial cells. Toxicology 223,<br />

181-190.<br />

Yip, L., Dart, R. C., and Gabow, P. A. (1994). Concepts and controversies in salicylate<br />

toxicity. Emerg Med Clin North Am 12, 351-364.<br />

Yonemitsu, K. (1986). Pharmacokinetic pr<strong>of</strong>ile <strong>of</strong> <strong>paraquat</strong> following intravenous<br />

administration to <strong>the</strong> rabbit. Forensic Sci Int 32, 33-42.<br />

Zweig, G., Shavit, N., and Avron, M. (1965). Diquat (I,I'-ethylene-2,2'-dipyridylium<br />

dibromide) in photo-reactions <strong>of</strong> isolated chloroplasts. Biochim Biophys Acta<br />

109, 332-346.<br />

249

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!